Sample records for additional energy bands

  1. Energy-banded ions in Saturn's magnetosphere

    NASA Astrophysics Data System (ADS)

    Thomsen, M. F.; Badman, S. V.; Jackman, C. M.; Jia, X.; Kivelson, M. G.; Kurth, W. S.

    2017-05-01

    Using data from the Cassini Plasma Spectrometer ion mass spectrometer, we report the first observation of energy-banded ions at Saturn. Observed near midnight at relatively high magnetic latitudes, the banded ions are dominantly H+, and they occupy the range of energies typically associated with the thermal pickup distribution in the inner magnetosphere (L < 10), but their energies decline monotonically with increasing radial distance (or time or decreasing latitude). Their pitch angle distribution suggests a source at low (or slightly southern) latitudes. The band energies, including their pitch angle dependence, are consistent with a bounce-resonant interaction between thermal H+ ions and the standing wave structure of a field line resonance. There is additional evidence in the pitch angle dependence of the band energies that the particles in each band may have a common time of flight from their most recent interaction with the wave, which may have been at slightly southern latitudes. Thus, while the particles are basically bounce resonant, their energization may be dominated by their most recent encounter with the standing wave.Plain Language SummaryDuring an outbound passage by the Cassini spacecraft through Saturn's inner magnetosphere, ion <span class="hlt">energy</span> distributions were observed that featured discrete flux peaks at regularly spaced <span class="hlt">energies</span>. The peaks persisted over several hours and several Saturn radii of distance away from the planet. We show that these "<span class="hlt">bands</span>" of ions are plausibly the result of an interaction between the Saturnian plasma and standing waves that form along the magnetospheric magnetic field lines. These observations are the first reported evidence that such standing waves may be present in the inner magnetosphere, where they could contribute to the radial transport of Saturn's radiation belt particles.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017ApSS..407...99K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017ApSS..407...99K"><span><span class="hlt">Energy</span> <span class="hlt">band</span> alignment of antiferroelectric (Pb,La)(Zr,Sn,Ti)O3</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Klein, Andreas; Lohaus, Christian; Reiser, Patrick; Dimesso, Lucangelo; Wang, Xiucai; Yang, Tongqing</p> <p>2017-06-01</p> <p>The <span class="hlt">energy</span> <span class="hlt">band</span> alignment of antiferroelectric (Pb,La)(Zr,Sn,Ti)O3 is studied with photoelectron spectroscopy using interfaces with high work function RuO2 and low work function Sn-doped In2O3 (ITO). It is demonstrated how spectral deconvolution can be used to determine absolute Schottky barrier heights for insulating materials with a high accuracy. Using this approach it is found that the valence <span class="hlt">band</span> maximum <span class="hlt">energy</span> of (Pb,La)(Zr,Sn,Ti)O3 is found to be comparable to that of Pb- and Bi-containing ferroelectric materials, which is ∼1 eV higher than that of BaTiO3. The results provide <span class="hlt">additional</span> evidence for the occupation of the 6s orbitals as origin of the higher valence <span class="hlt">band</span> maximum, which is directly related to the electrical properties of such compounds. The results also verify that the <span class="hlt">energy</span> <span class="hlt">band</span> alignment determined by photoelectron spectroscopy of as-deposited electrodes is not influenced by polarisation. The electronic structure of (Pb,La)(Zr,Sn,Ti)O3 should enable doping of the material without strongly modifying its insulating properties, which is crucial for high <span class="hlt">energy</span> density capacitors. Moreover, the position of the <span class="hlt">energy</span> <span class="hlt">bands</span> should result in a great freedom of selecting electrode materials in terms of avoiding charge injection.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018PhRvB..97m4521L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PhRvB..97m4521L"><span>Exotic superconductivity with enhanced <span class="hlt">energy</span> scales in materials with three <span class="hlt">band</span> crossings</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lin, Yu-Ping; Nandkishore, Rahul M.</p> <p>2018-04-01</p> <p>Three <span class="hlt">band</span> crossings can arise in three-dimensional quantum materials with certain space group symmetries. The low <span class="hlt">energy</span> Hamiltonian supports spin one fermions and a flat <span class="hlt">band</span>. We study the pairing problem in this setting. We write down a minimal BCS Hamiltonian and decompose it into spin-orbit coupled irreducible pairing channels. We then solve the resulting gap equations in channels with zero total angular momentum. We find that in the s-wave spin singlet channel (and also in an unusual d-wave `spin quintet' channel), superconductivity is enormously enhanced, with a possibility for the critical temperature to be linear in interaction strength. Meanwhile, in the p-wave spin triplet channel, the superconductivity exhibits features of conventional BCS theory due to the absence of flat <span class="hlt">band</span> pairing. Three <span class="hlt">band</span> crossings thus represent an exciting new platform for realizing exotic superconducting states with enhanced <span class="hlt">energy</span> scales. We also discuss the effects of doping, nonzero temperature, and of retaining <span class="hlt">additional</span> terms in the k .p expansion of the Hamiltonian.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2007PhRvL..99r6801Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2007PhRvL..99r6801Y"><span>Quasiparticle <span class="hlt">Energies</span> and <span class="hlt">Band</span> Gaps in Graphene Nanoribbons</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yang, Li; Park, Cheol-Hwan; Son, Young-Woo; Cohen, Marvin L.; Louie, Steven G.</p> <p>2007-11-01</p> <p>We present calculations of the quasiparticle <span class="hlt">energies</span> and <span class="hlt">band</span> gaps of graphene nanoribbons (GNRs) carried out using a first-principles many-electron Green’s function approach within the GW approximation. Because of the quasi-one-dimensional nature of a GNR, electron-electron interaction effects due to the enhanced screened Coulomb interaction and confinement geometry greatly influence the quasiparticle <span class="hlt">band</span> gap. Compared with previous tight-binding and density functional theory studies, our calculated quasiparticle <span class="hlt">band</span> gaps show significant self-<span class="hlt">energy</span> corrections for both armchair and zigzag GNRs, in the range of 0.5 3.0 eV for ribbons of width 2.4 0.4 nm. The quasiparticle <span class="hlt">band</span> gaps found here suggest that use of GNRs for electronic device components in ambient conditions may be viable.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26323493','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26323493"><span>Quantitative analysis on electric dipole <span class="hlt">energy</span> in Rashba <span class="hlt">band</span> splitting.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Hong, Jisook; Rhim, Jun-Won; Kim, Changyoung; Ryong Park, Seung; Hoon Shim, Ji</p> <p>2015-09-01</p> <p>We report on quantitative comparison between the electric dipole <span class="hlt">energy</span> and the Rashba <span class="hlt">band</span> splitting in model systems of Bi and Sb triangular monolayers under a perpendicular electric field. We used both first-principles and tight binding calculations on p-orbitals with spin-orbit coupling. First-principles calculation shows Rashba <span class="hlt">band</span> splitting in both systems. It also shows asymmetric charge distributions in the Rashba split <span class="hlt">bands</span> which are induced by the orbital angular momentum. We calculated the electric dipole <span class="hlt">energies</span> from coupling of the asymmetric charge distribution and external electric field, and compared it to the Rashba splitting. Remarkably, the total split <span class="hlt">energy</span> is found to come mostly from the difference in the electric dipole <span class="hlt">energy</span> for both Bi and Sb systems. A perturbative approach for long wave length limit starting from tight binding calculation also supports that the Rashba <span class="hlt">band</span> splitting originates mostly from the electric dipole <span class="hlt">energy</span> difference in the strong atomic spin-orbit coupling regime.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4555038','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4555038"><span>Quantitative analysis on electric dipole <span class="hlt">energy</span> in Rashba <span class="hlt">band</span> splitting</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Hong, Jisook; Rhim, Jun-Won; Kim, Changyoung; Ryong Park, Seung; Hoon Shim, Ji</p> <p>2015-01-01</p> <p>We report on quantitative comparison between the electric dipole <span class="hlt">energy</span> and the Rashba <span class="hlt">band</span> splitting in model systems of Bi and Sb triangular monolayers under a perpendicular electric field. We used both first-principles and tight binding calculations on p-orbitals with spin-orbit coupling. First-principles calculation shows Rashba <span class="hlt">band</span> splitting in both systems. It also shows asymmetric charge distributions in the Rashba split <span class="hlt">bands</span> which are induced by the orbital angular momentum. We calculated the electric dipole <span class="hlt">energies</span> from coupling of the asymmetric charge distribution and external electric field, and compared it to the Rashba splitting. Remarkably, the total split <span class="hlt">energy</span> is found to come mostly from the difference in the electric dipole <span class="hlt">energy</span> for both Bi and Sb systems. A perturbative approach for long wave length limit starting from tight binding calculation also supports that the Rashba <span class="hlt">band</span> splitting originates mostly from the electric dipole <span class="hlt">energy</span> difference in the strong atomic spin-orbit coupling regime. PMID:26323493</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2009SPIE.7504E..0GG','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009SPIE.7504E..0GG"><span>Ultrafast laser-induced modifications of <span class="hlt">energy</span> <span class="hlt">bands</span> of non-metal crystals</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gruzdev, Vitaly</p> <p>2009-10-01</p> <p>Ultrafast laser-induced variations of electron <span class="hlt">energy</span> <span class="hlt">bands</span> of transparent solids significantly influence ionization and conduction-<span class="hlt">band</span> electron absorption driving the initial stage of laser-induced damage (LID). The mechanisms of the variations are attributed to changing electron functions from bonding to anti-bonding configuration via laser-induced ionization; laser-driven electron oscillations in quasi-momentum space; and direct distortion of the inter-atomic potential by electric field of laser radiation. The ionization results in the <span class="hlt">band</span>-structure modification via accumulation of broken chemical bonds between atoms and provides significant contribution to the overall modification only when enough excited electrons are accumulated in the conduction <span class="hlt">band</span>. The oscillations are associated with modification of electron <span class="hlt">energy</span> by pondermotive potential of the oscillations. The direct action of radiation's electric field leads to specific high-frequency Franz-Keldysh effect (FKE) spreading the allowed electron states into the <span class="hlt">bands</span> of forbidden <span class="hlt">energy</span>. Those processes determine the effective <span class="hlt">band</span> gap that is a laser-driven <span class="hlt">energy</span> gap between the modified electron <span class="hlt">energy</span> <span class="hlt">bands</span>. Among those mechanisms, the latter two provide reversible <span class="hlt">band</span>-structure modification that takes place from the beginning of the ionization and are, therefore, of special interest due to their strong influence on the initial stage of the ionization. The pondermotive potential results either in monotonous increase or oscillatory variations of the effective <span class="hlt">band</span> gap that has been taken into account in some ionization models. The classical FKE provides decrease of the <span class="hlt">band</span> gap. We analyzing the competition between those two opposite trends of the effective-<span class="hlt">band</span>-gap variations and discuss applications of those effects for considerations of the laser-induced damage and its threshold in transparent solids.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010AIPC.1278..131G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010AIPC.1278..131G"><span>Laser-Induced Modification Of <span class="hlt">Energy</span> <span class="hlt">Bands</span> Of Transparent Solids</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gruzdev, Vitaly</p> <p>2010-10-01</p> <p>Laser-induced variations of electron <span class="hlt">energy</span> <span class="hlt">bands</span> of transparent solids significantly affect the initial stages of laser-induced ablation (LIA) influencing rates of ionization and light absorption by conduction-<span class="hlt">band</span> electrons. We analyze fast variations with characteristic duration in femto-second time domain that include: 1) switching electron functions from bonding to anti-bonding configuration due to laser-induced ionization; 2) laser-driven oscillations of electrons in quasi-momentum space; and 3) direct distortion of the inter-atomic potential by electric field of laser radiation. Among those effects, the latter two have zero delay and reversibly modify <span class="hlt">band</span> structure taking place from the beginning of laser action. They are of special interest due to their strong influence on the initial stage and threshold of laser ablation. The oscillations modify the electron-<span class="hlt">energy</span> <span class="hlt">bands</span> by adding pondermotive potential. The direct action of radiation's electric field leads to high-frequency Franz-Keldysh effect (FKE) spreading the allowed electron states into the forbidden-<span class="hlt">energy</span> <span class="hlt">bands</span>. FKE provides decrease of the effective <span class="hlt">band</span> gap while the electron oscillations lead either to monotonous increase or oscillatory variations of the gap. We analyze the competition between those two opposite trends and their role in initiating LIA.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28325035','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28325035"><span>Calculation of <span class="hlt">Energy</span> Diagram of Asymmetric Graded-<span class="hlt">Band</span>-Gap Semiconductor Superlattices.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Monastyrskii, Liubomyr S; Sokolovskii, Bogdan S; Alekseichyk, Mariya P</p> <p>2017-12-01</p> <p>The paper theoretically investigates the peculiarities of <span class="hlt">energy</span> diagram of asymmetric graded-<span class="hlt">band</span>-gap superlattices with linear coordinate dependences of <span class="hlt">band</span> gap and electron affinity. For calculating the <span class="hlt">energy</span> diagram of asymmetric graded-<span class="hlt">band</span>-gap superlattices, linearized Poisson's equation has been solved for the two layers forming a period of the superlattice. The obtained coordinate dependences of edges of the conduction and valence <span class="hlt">bands</span> demonstrate substantial transformation of the shape of the <span class="hlt">energy</span> diagram at changing the period of the lattice and the ratio of width of the adjacent layers. The most marked changes in the <span class="hlt">energy</span> diagram take place when the period of lattice is comparable with the Debye screening length. In the case when the lattice period is much smaller that the Debye screening length, the <span class="hlt">energy</span> diagram has the shape of a sawtooth-like pattern.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017PhRvB..96o5439K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017PhRvB..96o5439K"><span>Quasiparticle <span class="hlt">energy</span> <span class="hlt">bands</span> and Fermi surfaces of monolayer NbSe2</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kim, Sejoong; Son, Young-Woo</p> <p>2017-10-01</p> <p>A quasiparticle <span class="hlt">band</span> structure of a single layer 2 H -NbSe2 is reported by using first-principles G W calculation. We show that a self-<span class="hlt">energy</span> correction increases the width of a partially occupied <span class="hlt">band</span> and alters its Fermi surface shape when comparing those using conventional mean-field calculation methods. Owing to a broken inversion symmetry in the trigonal prismatic single layer structure, the spin-orbit interaction is included and its impact on the Fermi surface and quasiparticle <span class="hlt">energy</span> <span class="hlt">bands</span> are discussed. We also calculate the doping dependent static susceptibilities from the <span class="hlt">band</span> structures obtained by the mean-field calculation as well as G W calculation with and without spin-orbit interactions. A complete tight-binding model is constructed within the three-<span class="hlt">band</span> third nearest neighbor hoppings and is shown to reproduce our G W quasiparticle <span class="hlt">energy</span> <span class="hlt">bands</span> and Fermi surface very well. Considering variations of the Fermi surface shapes depending on self-<span class="hlt">energy</span> corrections and spin-orbit interactions, we discuss the formations of charge density wave (CDW) with different dielectric environments and their implications on recent controversial experimental results on CDW transition temperatures.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1999PhRvB..5910119X','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1999PhRvB..5910119X"><span><span class="hlt">Energy</span> <span class="hlt">bands</span> and acceptor binding <span class="hlt">energies</span> of GaN</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Xia, Jian-Bai; Cheah, K. W.; Wang, Xiao-Liang; Sun, Dian-Zhao; Kong, Mei-Ying</p> <p>1999-04-01</p> <p>The <span class="hlt">energy</span> <span class="hlt">bands</span> of zinc-blende and wurtzite GaN are calculated with the empirical pseudopotential method, and the pseudopotential parameters for Ga and N atoms are given. The calculated <span class="hlt">energy</span> <span class="hlt">bands</span> are in agreement with those obtained by the ab initio method. The effective-mass theory for the semiconductors of wurtzite structure is established, and the effective-mass parameters of GaN for both structures are given. The binding <span class="hlt">energies</span> of acceptor states are calculated by solving strictly the effective-mass equations. The binding <span class="hlt">energies</span> of donor and acceptor are 24 and 142 meV for the zinc-blende structure, 20 and 131, and 97 meV for the wurtzite structure, respectively, which are consistent with recent experimental results. It is proposed that there are two kinds of acceptor in wurtzite GaN. One kind is the general acceptor such as C, which substitutes N, which satisfies the effective-mass theory. The other kind of acceptor includes Mg, Zn, Cd, etc., the binding <span class="hlt">energy</span> of these acceptors is deviated from that given by the effective-mass theory. In this report, wurtzite GaN is grown by the molecular-beam epitaxy method, and the photoluminescence spectra were measured. Three main peaks are assigned to the donor-acceptor transitions from two kinds of acceptors. Some of the transitions were identified as coming from the cubic phase of GaN, which appears randomly within the predominantly hexagonal material.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22489482-tensile-strain-effect-inducing-indirect-direct-band-gap-transition-reducing-band-gap-energy-ge','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22489482-tensile-strain-effect-inducing-indirect-direct-band-gap-transition-reducing-band-gap-energy-ge"><span>Tensile-strain effect of inducing the indirect-to-direct <span class="hlt">band</span>-gap transition and reducing the <span class="hlt">band</span>-gap <span class="hlt">energy</span> of Ge</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Inaoka, Takeshi, E-mail: inaoka@phys.u-ryukyu.ac.jp; Furukawa, Takuro; Toma, Ryo</p> <p></p> <p>By means of a hybrid density-functional method, we investigate the tensile-strain effect of inducing the indirect-to-direct <span class="hlt">band</span>-gap transition and reducing the <span class="hlt">band</span>-gap <span class="hlt">energy</span> of Ge. We consider [001], [111], and [110] uniaxial tensility and (001), (111), and (110) biaxial tensility. Under the condition of no normal stress, we determine both normal compression and internal strain, namely, relative displacement of two atoms in the primitive unit cell, by minimizing the total <span class="hlt">energy</span>. We identify those strain types which can induce the <span class="hlt">band</span>-gap transition, and evaluate the critical strain coefficient where the gap transition occurs. Either normal compression or internal strain operatesmore » unfavorably to induce the gap transition, which raises the critical strain coefficient or even blocks the transition. We also examine how each type of tensile strain decreases the <span class="hlt">band</span>-gap <span class="hlt">energy</span>, depending on its orientation. Our analysis clearly shows that synergistic operation of strain orientation and <span class="hlt">band</span> anisotropy has a great influence on the gap transition and the gap <span class="hlt">energy</span>.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2005ApPhL..87c2102K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2005ApPhL..87c2102K"><span><span class="hlt">Band</span> gap and <span class="hlt">band</span> offset of (GaIn)(PSb) lattice matched to InP</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Köhler, F.; Böhm, G.; Meyer, R.; Amann, M.-C.</p> <p>2005-07-01</p> <p>Metastable (GaxIn1-x)(PySb1-y) layers were grown on (001) InP substrates by gas source molecular beam epitaxy. Low-temperature photoluminescence spectroscopy was applied to these heterostructures and revealed spatially indirect <span class="hlt">band-to-band</span> recombination of electrons localized in the InP with holes in the (GaxIn1-x)(PySb1-y). In <span class="hlt">addition</span>, samples with layer thicknesses larger than 100nm showed direct PL across the <span class="hlt">band</span> gap of (GaxIn1-x)(PySb1-y). <span class="hlt">Band</span>-gap <span class="hlt">energies</span> and <span class="hlt">band</span> offset <span class="hlt">energies</span> of (GaxIn1-x)(PySb1-y) relative to InP were derived from these PL data. A strong bowing parameter was observed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014CoPhC.185.1195S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014CoPhC.185.1195S"><span>Improved cache performance in Monte Carlo transport calculations using <span class="hlt">energy</span> <span class="hlt">banding</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Siegel, A.; Smith, K.; Felker, K.; Romano, P.; Forget, B.; Beckman, P.</p> <p>2014-04-01</p> <p>We present an <span class="hlt">energy</span> <span class="hlt">banding</span> algorithm for Monte Carlo (MC) neutral particle transport simulations which depend on large cross section lookup tables. In MC codes, read-only cross section data tables are accessed frequently, exhibit poor locality, and are typically too much large to fit in fast memory. Thus, performance is often limited by long latencies to RAM, or by off-node communication latencies when the data footprint is very large and must be decomposed on a distributed memory machine. The proposed <span class="hlt">energy</span> <span class="hlt">banding</span> algorithm allows maximal temporal reuse of data in <span class="hlt">band</span> sizes that can flexibly accommodate different architectural features. The <span class="hlt">energy</span> <span class="hlt">banding</span> algorithm is general and has a number of benefits compared to the traditional approach. In the present analysis we explore its potential to achieve improvements in time-to-solution on modern cache-based architectures.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JPhD...50NLT02G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JPhD...50NLT02G"><span>A novel theoretical model for the temperature dependence of <span class="hlt">band</span> gap <span class="hlt">energy</span> in semiconductors</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Geng, Peiji; Li, Weiguo; Zhang, Xianhe; Zhang, Xuyao; Deng, Yong; Kou, Haibo</p> <p>2017-10-01</p> <p>We report a novel theoretical model without any fitting parameters for the temperature dependence of <span class="hlt">band</span> gap <span class="hlt">energy</span> in semiconductors. This model relates the <span class="hlt">band</span> gap <span class="hlt">energy</span> at the elevated temperature to that at the arbitrary reference temperature. As examples, the <span class="hlt">band</span> gap <span class="hlt">energies</span> of Si, Ge, AlN, GaN, InP, InAs, ZnO, ZnS, ZnSe and GaAs at temperatures below 400 K are calculated and are in good agreement with the experimental results. Meanwhile, the <span class="hlt">band</span> gap <span class="hlt">energies</span> at high temperatures (T  >  400 K) are predicted, which are greater than the experimental results, and the reasonable analysis is carried out as well. Under low temperatures, the effect of lattice expansion on the <span class="hlt">band</span> gap <span class="hlt">energy</span> is very small, but it has much influence on the <span class="hlt">band</span> gap <span class="hlt">energy</span> at high temperatures. Therefore, it is necessary to consider the effect of lattice expansion at high temperatures, and the method considering the effect of lattice expansion has also been given. The model has distinct advantages compared with the widely quoted Varshni’s semi-empirical equation from the aspect of modeling, physical meaning and application. The study provides a convenient method to determine the <span class="hlt">band</span> gap <span class="hlt">energy</span> under different temperatures.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2004PhDT........31H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2004PhDT........31H"><span>Spectroscopic study of hafnium silicate alloys prepared by RPECVD: Comparisons between conduction/valence <span class="hlt">band</span> offset <span class="hlt">energies</span> and optical <span class="hlt">band</span> gaps</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hong, Joon Goo</p> <p></p> <p>Aggressive scaling of devices has continued to improve MOSFET transistor performance. As lateral device dimensions continue to decrease, gate oxide thickness must be scaled down. As one of the promising high k alternative gate oxide materials, HfO2 and its silicates were investigated to understand their direct tunneling behavior by studying <span class="hlt">band</span> offset <span class="hlt">energies</span> with spectroscopy and electrical characterization. Local bonding change of remote plasma deposited (HfO2)x(SiO 2)1-x alloys were characterized by Fourier transform infrared (FTIR) spectroscopy, x-ray photoelectron spectroscopy (XPS), and Auger electron spectroscopy (AES) as a function of alloy composition, x. Two different precursors with Hf Nitrato and Hf-tert-butoxide were tested to have amorphous deposition. Film composition was determined off-line by Rutherford backscattering spectroscopy (RBS) and these results were calibrated with on-line AES. As deposited Hf-silicate alloys were characterized by off-line XPS and AES for their chemical shifts interpreting with a partial charge transfer model as well as coordination changes. Sigmoidal dependence of valence <span class="hlt">band</span> offset <span class="hlt">energies</span> was observed. Hf 5d* state is fixed at the bottom of the conduction <span class="hlt">band</span> and located at 1.3 +/- 0.2 eV above the top of the Si conduction <span class="hlt">band</span> as a conduction <span class="hlt">band</span> offset by x-ray absorption spectroscopy (XAS). Optical <span class="hlt">band</span> gap <span class="hlt">energy</span> changes were observed with vacuum ultra violet spectroscopic ellipsometry (VUVSE) to verify compositional dependence of conduction and valence <span class="hlt">band</span> offset <span class="hlt">energy</span> changes. 1 nm EOT normalized tunneling current with Wentzel-Kramer-Brillouin (WKB) simulation based on the <span class="hlt">band</span> offset study and Franz two <span class="hlt">band</span> model showed the minimum at the intermediate composition matching with the experimental data. Non-linear trend in tunneling current was observed because the increases in physical thickness were mitigated by reductions in <span class="hlt">band</span> offset <span class="hlt">energies</span> and effective mass for tunneling. C-V curves were compared</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015JSemi..36a3001A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015JSemi..36a3001A"><span>The calculation of <span class="hlt">band</span> gap <span class="hlt">energy</span> in zinc oxide films</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Arif, Ali; Belahssen, Okba; Gareh, Salim; Benramache, Said</p> <p>2015-01-01</p> <p>We investigated the optical properties of undoped zinc oxide thin films as the n-type semiconductor; the thin films were deposited at different precursor molarities by ultrasonic spray and spray pyrolysis techniques. The thin films were deposited at different substrate temperatures ranging between 200 and 500 °C. In this paper, we present a new approach to control the optical gap <span class="hlt">energy</span> of ZnO thin films by concentration of the ZnO solution and substrate temperatures from experimental data, which were published in international journals. The model proposed to calculate the <span class="hlt">band</span> gap <span class="hlt">energy</span> with the Urbach <span class="hlt">energy</span> was investigated. The relation between the experimental data and theoretical calculation suggests that the <span class="hlt">band</span> gap <span class="hlt">energies</span> are predominantly estimated by the Urbach <span class="hlt">energies</span>, film transparency, and concentration of the ZnO solution and substrate temperatures. The measurements by these proposal models are in qualitative agreements with the experimental data; the correlation coefficient values were varied in the range 0.96-0.99999, indicating high quality representation of data based on Equation (2), so that the relative errors of all calculation are smaller than 4%. Thus, one can suppose that the undoped ZnO thin films are chemically purer and have many fewer defects and less disorder owing to an almost complete chemical decomposition and contained higher optical <span class="hlt">band</span> gap <span class="hlt">energy</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25247447','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25247447"><span><span class="hlt">Energy</span> <span class="hlt">band</span> gap and optical transition of metal ion modified double crossover DNA lattices.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Dugasani, Sreekantha Reddy; Ha, Taewoo; Gnapareddy, Bramaramba; Choi, Kyujin; Lee, Junwye; Kim, Byeonghoon; Kim, Jae Hoon; Park, Sung Ha</p> <p>2014-10-22</p> <p>We report on the <span class="hlt">energy</span> <span class="hlt">band</span> gap and optical transition of a series of divalent metal ion (Cu(2+), Ni(2+), Zn(2+), and Co(2+)) modified DNA (M-DNA) double crossover (DX) lattices fabricated on fused silica by the substrate-assisted growth (SAG) method. We demonstrate how the degree of coverage of the DX lattices is influenced by the DX monomer concentration and also analyze the <span class="hlt">band</span> gaps of the M-DNA lattices. The <span class="hlt">energy</span> <span class="hlt">band</span> gap of the M-DNA, between the lowest unoccupied molecular orbital (LUMO) and the highest occupied molecular orbital (HOMO), ranges from 4.67 to 4.98 eV as judged by optical transitions. Relative to the <span class="hlt">band</span> gap of a pristine DNA molecule (4.69 eV), the <span class="hlt">band</span> gap of the M-DNA lattices increases with metal ion doping up to a critical concentration and then decreases with further doping. Interestingly, except for the case of Ni(2+), the onset of the second absorption <span class="hlt">band</span> shifts to a lower <span class="hlt">energy</span> until a critical concentration and then shifts to a higher <span class="hlt">energy</span> with further increasing the metal ion concentration, which is consistent with the evolution of electrical transport characteristics. Our results show that controllable metal ion doping is an effective method to tune the <span class="hlt">band</span> gap <span class="hlt">energy</span> of DNA-based nanostructures.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016OptMa..53..134K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016OptMa..53..134K"><span>Effects of optical <span class="hlt">band</span> gap <span class="hlt">energy</span>, <span class="hlt">band</span> tail <span class="hlt">energy</span> and particle shape on photocatalytic activities of different ZnO nanostructures prepared by a hydrothermal method</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Klubnuan, Sarunya; Suwanboon, Sumetha; Amornpitoksuk, Pongsaton</p> <p>2016-03-01</p> <p>The dependence of the crystallite size and the <span class="hlt">band</span> tail <span class="hlt">energy</span> on the optical properties, particle shape and oxygen vacancy of different ZnO nanostructures to catalyse photocatalytic degradation was investigated. The ZnO nanoplatelets and mesh-like ZnO lamellae were synthesized from the PEO19-b-PPO3 modified zinc acetate dihydrate using aqueous KOH and CO(NH2)2 solutions, respectively via a hydrothermal method. The <span class="hlt">band</span> tail <span class="hlt">energy</span> of the ZnO nanostructures had more influence on the <span class="hlt">band</span> gap <span class="hlt">energy</span> than the crystallite size. The photocatalytic degradation of methylene blue increased as a function of the irradiation time, the amount of oxygen vacancy and the intensity of the (0 0 0 2) plane. The ZnO nanoplatelets exhibited a better photocatalytic degradation of methylene blue than the mesh-like ZnO lamellae due to the migration of the photoelectrons and holes to the (0 0 0 1) and (0 0 0 -1) planes, respectively under the internal electric field, that resulted in the enhancement of the photocatalytic activities.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013JPCS...74...45S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013JPCS...74...45S"><span>A simplified approach to the <span class="hlt">band</span> gap correction of defect formation <span class="hlt">energies</span>: Al, Ga, and In-doped ZnO</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Saniz, R.; Xu, Y.; Matsubara, M.; Amini, M. N.; Dixit, H.; Lamoen, D.; Partoens, B.</p> <p>2013-01-01</p> <p>The calculation of defect levels in semiconductors within a density functional theory approach suffers greatly from the <span class="hlt">band</span> gap problem. We propose a <span class="hlt">band</span> gap correction scheme that is based on the separation of <span class="hlt">energy</span> differences in electron <span class="hlt">addition</span> and relaxation <span class="hlt">energies</span>. We show that it can predict defect levels with a reasonable accuracy, particularly in the case of defects with conduction <span class="hlt">band</span> character, and yet is simple and computationally economical. We apply this method to ZnO doped with group III elements (Al, Ga, In). As expected from experiment, the results indicate that Zn substitutional doping is preferred over interstitial doping in Al, Ga, and In-doped ZnO, under both zinc-rich and oxygen-rich conditions. Further, all three dopants act as shallow donors, with the +1 charge state having the most advantageous formation <span class="hlt">energy</span>. Also, doping effects on the electronic structure of ZnO are sufficiently mild so as to affect little the fundamental <span class="hlt">band</span> gap and lowest conduction <span class="hlt">bands</span> dispersion, which secures their n-type transparent conducting behavior. A comparison with the extrapolation method based on LDA+U calculations and with the Heyd-Scuseria-Ernzerhof hybrid functional (HSE) shows the reliability of the proposed scheme in predicting the thermodynamic transition levels in shallow donor systems.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li class="active"><span>1</span></li> <li><a href="#" onclick='return showDiv("page_2");'>2</a></li> <li><a href="#" onclick='return showDiv("page_3");'>3</a></li> <li><a href="#" onclick='return showDiv("page_4");'>4</a></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_1 --> <div id="page_2" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_1");'>1</a></li> <li class="active"><span>2</span></li> <li><a href="#" onclick='return showDiv("page_3");'>3</a></li> <li><a href="#" onclick='return showDiv("page_4");'>4</a></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="21"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://eric.ed.gov/?q=%22light-emitting+diode%22+OR+lighting&pg=3&id=EJ829407','ERIC'); return false;" href="https://eric.ed.gov/?q=%22light-emitting+diode%22+OR+lighting&pg=3&id=EJ829407"><span>Simple Experimental Verification of the Relation between the <span class="hlt">Band</span>-Gap <span class="hlt">Energy</span> and the <span class="hlt">Energy</span> of Photons Emitted by LEDs</span></a></p> <p><a target="_blank" href="http://www.eric.ed.gov/ERICWebPortal/search/extended.jsp?_pageLabel=advanced">ERIC Educational Resources Information Center</a></p> <p>Precker, Jurgen W.</p> <p>2007-01-01</p> <p>The wavelength of the light emitted by a light-emitting diode (LED) is intimately related to the <span class="hlt">band</span>-gap <span class="hlt">energy</span> of the semiconductor from which the LED is made. We experimentally estimate the <span class="hlt">band</span>-gap <span class="hlt">energies</span> of several types of LEDs, and compare them with the <span class="hlt">energies</span> of the emitted light, which ranges from infrared to white. In spite of…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015MNRAS.452.3666S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015MNRAS.452.3666S"><span><span class="hlt">Energy</span> dependence of the <span class="hlt">band</span>-limited noise in black hole X-ray binaries★</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Stiele, H.; Yu, W.</p> <p>2015-10-01</p> <p>Black hole low-mass X-ray binaries show a variety of variability features, which manifest as narrow peak-like structures superposed on broad noise components in power density spectra in the hard X-ray emission. In this work, we study variability properties of the <span class="hlt">band</span>-limited noise component during the low-hard state for a sample of black hole X-ray binaries. We investigate the characteristic frequency and amplitude of the <span class="hlt">band</span>-limited noise component and study covariance spectra. For observations that show a noise component with a characteristic frequency above 1 Hz in the hard <span class="hlt">energy</span> <span class="hlt">band</span> (4-8 keV), we found this very same component at a lower frequency in the soft <span class="hlt">band</span> (1-2 keV). This difference in characteristic frequency is an indication that while both the soft and the hard <span class="hlt">band</span> photons contribute to the same <span class="hlt">band</span>-limited noise component, which likely represents the modulation of the mass accretion rate, the origin of the soft photons is actually further away from the black hole than the hard photons. Thus, the soft photons are characterized by larger radii, lower frequencies and softer <span class="hlt">energies</span>, and are probably associated with a smaller optical depth for Comptonization up-scattering from the outer layer of the corona, or suggest a temperature gradient of the corona. We interpret this <span class="hlt">energy</span> dependence within the picture of <span class="hlt">energy</span>-dependent power density states as a hint that the contribution of the up-scattered photons originating in the outskirts of the Comptonizing corona to the overall emission in the soft <span class="hlt">band</span> is becoming significant.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22261866-theoretical-study-energy-states-two-dimensional-electron-gas-pseudomorphically-strained-inas-hemts-taking-account-non-parabolicity-conduction-band','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22261866-theoretical-study-energy-states-two-dimensional-electron-gas-pseudomorphically-strained-inas-hemts-taking-account-non-parabolicity-conduction-band"><span>Theoretical study of <span class="hlt">energy</span> states of two-dimensional electron gas in pseudomorphically strained InAs HEMTs taking into account the non-parabolicity of the conduction <span class="hlt">band</span></span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Nishio, Yui; Yamaguchi, Satoshi; Yamazaki, Youichi</p> <p>2013-12-04</p> <p>We determined rigorously the <span class="hlt">energy</span> states of a two-dimensional electron gas (2DEG) in high electron mobility transistors (HEMTs) with a pseudomorphically strained InAs channel (InAs PHEMTs) taking into account the non-parabolicity of the conduction <span class="hlt">band</span> for InAs. The sheet carrier concentration of 2DEG for the non-parabolic <span class="hlt">energy</span> <span class="hlt">band</span> was about 50% larger than that for the parabolic <span class="hlt">energy</span> <span class="hlt">band</span> and most of the electrons are confined strongly in the InAs layer. In <span class="hlt">addition</span>, the threshold voltage for InAs PHEMTs was about 0.21 V lower than that for conventional InGaAs HEMTs.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29578678','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29578678"><span><span class="hlt">Band</span> Edge Dynamics and Multiexciton Generation in Narrow <span class="hlt">Band</span> Gap HgTe Nanocrystals.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Livache, Clément; Goubet, Nicolas; Martinez, Bertille; Jagtap, Amardeep; Qu, Junling; Ithurria, Sandrine; Silly, Mathieu G; Dubertret, Benoit; Lhuillier, Emmanuel</p> <p>2018-04-11</p> <p>Mercury chalcogenide nanocrystals and especially HgTe appear as an interesting platform for the design of low cost mid-infrared (mid-IR) detectors. Nevertheless, their electronic structure and transport properties remain poorly understood, and some critical aspects such as the carrier relaxation dynamics at the <span class="hlt">band</span> edge have been pushed under the rug. Some of the previous reports on dynamics are setup-limited, and all of them have been obtained using photon <span class="hlt">energy</span> far above the <span class="hlt">band</span> edge. These observations raise two main questions: (i) what are the carrier dynamics at the <span class="hlt">band</span> edge and (ii) should we expect some <span class="hlt">additional</span> effect (multiexciton generation (MEG)) as such narrow <span class="hlt">band</span> gap materials are excited far above the <span class="hlt">band</span> edge? To answer these questions, we developed a high-bandwidth setup that allows us to understand and compare the carrier dynamics resonantly pumped at the <span class="hlt">band</span> edge in the mid-IR and far above the <span class="hlt">band</span> edge. We demonstrate that fast (>50 MHz) photoresponse can be obtained even in the mid-IR and that MEG is occurring in HgTe nanocrystal arrays with a threshold around 3 times the <span class="hlt">band</span> edge <span class="hlt">energy</span>. Furthermore, the photoresponse can be effectively tuned in magnitude and sign using a phototransistor configuration.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1999SPIE.3797..178S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1999SPIE.3797..178S"><span><span class="hlt">Energy</span> level alignment and <span class="hlt">band</span> bending at organic interfaces</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Seki, Kazuhiko; Oji, Hiroshi; Ito, Eisuke; Hayashi, Naoki; Ouchi, Yukio; Ishii, Hisao</p> <p>1999-12-01</p> <p>Recent progress in the study of the <span class="hlt">energy</span> level alignment and <span class="hlt">band</span> bending at organic interfaces is reviewed, taking the examples mainly from the results of the group of the authors using ultraviolet photoelectron spectroscopy (UPS), metastable atom electron spectroscopy (MAES), and Kelvin probe method (KPM). As for the <span class="hlt">energy</span> level alignment right at the interface, the formation of an electric dipole layer is observed for most of the organic/metal interfaces, even when no significant chemical interaction is observed. The origin of this dipole layer is examined by accumulating the data of various combinations of organics and metals, and the results indicate combined contribution from (1) charge transfer (CT) between the organic molecule and the metal, and (2) pushback of the electrons spilled out from metal surface, for the case of nonpolar organic molecule physisorbed on metals. Other factors such as chemical interaction and the orientation of polar molecules are also pointed out. As for the <span class="hlt">band</span> bending, the careful examination of the existence/absence of <span class="hlt">band</span> bending of purified TPD* molecule deposited on various metals in ultrahigh vacuum (UHV) revealed negligible <span class="hlt">band</span> bending up to 100 nm thickness, and also the failure of the establishment of Fermi level alignment between organic layer and the metals. The implications of these findings are discussed, in relation to the future prospects of the studies in this field. (*:N,N'- diphenyl-N,N'-(3-methylphenyl)-1,1'-biphenyl-4,4'-diamine).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018npjQM...3....1T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018npjQM...3....1T"><span>Observation of Dirac-like <span class="hlt">energy</span> <span class="hlt">band</span> and ring-torus Fermi surface associated with the nodal line in topological insulator CaAgAs</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Takane, Daichi; Nakayama, Kosuke; Souma, Seigo; Wada, Taichi; Okamoto, Yoshihiko; Takenaka, Koshi; Yamakawa, Youichi; Yamakage, Ai; Mitsuhashi, Taichi; Horiba, Koji; Kumigashira, Hiroshi; Takahashi, Takashi; Sato, Takafumi</p> <p>2018-01-01</p> <p>One of key challenges in current material research is to search for new topological materials with inverted bulk-<span class="hlt">band</span> structure. In topological insulators, the <span class="hlt">band</span> inversion caused by strong spin-orbit coupling leads to opening of a <span class="hlt">band</span> gap in the entire Brillouin zone, whereas an <span class="hlt">additional</span> crystal symmetry such as point-group and nonsymmorphic symmetries sometimes prohibits the gap opening at/on specific points or line in momentum space, giving rise to topological semimetals. Despite many theoretical predictions of topological insulators/semimetals associated with such crystal symmetries, the experimental realization is still relatively scarce. Here, using angle-resolved photoemission spectroscopy with bulk-sensitive soft-x-ray photons, we experimentally demonstrate that hexagonal pnictide CaAgAs belongs to a new family of topological insulators characterized by the inverted <span class="hlt">band</span> structure and the mirror reflection symmetry of crystal. We have established the bulk valence-<span class="hlt">band</span> structure in three-dimensional Brillouin zone, and observed the Dirac-like <span class="hlt">energy</span> <span class="hlt">band</span> and ring-torus Fermi surface associated with the line node, where bulk valence and conducting <span class="hlt">bands</span> cross on a line in the momentum space under negligible spin-orbit coupling. Intriguingly, we found that no other <span class="hlt">bands</span> cross the Fermi level and therefore the low-<span class="hlt">energy</span> excitations are solely characterized by the Dirac-like <span class="hlt">band</span>. CaAgAs provides an excellent platform to study the interplay among low-<span class="hlt">energy</span> electron dynamics, crystal symmetry, and exotic topological properties.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/1411617-band-gaps-elastic-wave-propagation-periodic-composite-beam-structure-incorporating-microstructure-surface-energy-effects','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/1411617-band-gaps-elastic-wave-propagation-periodic-composite-beam-structure-incorporating-microstructure-surface-energy-effects"><span><span class="hlt">Band</span> Gaps for Elastic Wave Propagation in a Periodic Composite Beam Structure Incorporating Microstructure and Surface <span class="hlt">Energy</span> Effects</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Zhang, G. Y.; Gao, X. -L.; Bishop, J. E.</p> <p></p> <p>Here, a new model for determining <span class="hlt">band</span> gaps for elastic wave propagation in a periodic composite beam structure is developed using a non-classical Bernoulli–Euler beam model that incorporates the microstructure, surface <span class="hlt">energy</span> and rotational inertia effects. The Bloch theorem and transfer matrix method for periodic structures are employed in the formulation. The new model reduces to the classical elasticity-based model when both the microstructure and surface <span class="hlt">energy</span> effects are not considered. The <span class="hlt">band</span> gaps predicted by the new model depend on the microstructure and surface elasticity of each constituent material, the unit cell size, the rotational inertia, and the volumemore » fraction. To quantitatively illustrate the effects of these factors, a parametric study is conducted. The numerical results reveal that the <span class="hlt">band</span> gap predicted by the current non-classical model is always larger than that predicted by the classical model when the beam thickness is very small, but the difference is diminishing as the thickness becomes large. Also, it is found that the first frequency for producing the <span class="hlt">band</span> gap and the <span class="hlt">band</span> gap size decrease with the increase of the unit cell length according to both the current and classical models. In <span class="hlt">addition</span>, it is observed that the effect of the rotational inertia is larger when the exciting frequency is higher and the unit cell length is smaller. Furthermore, it is seen that the volume fraction has a significant effect on the <span class="hlt">band</span> gap size, and large <span class="hlt">band</span> gaps can be obtained by tailoring the volume fraction and material parameters.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1411617-band-gaps-elastic-wave-propagation-periodic-composite-beam-structure-incorporating-microstructure-surface-energy-effects','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1411617-band-gaps-elastic-wave-propagation-periodic-composite-beam-structure-incorporating-microstructure-surface-energy-effects"><span><span class="hlt">Band</span> Gaps for Elastic Wave Propagation in a Periodic Composite Beam Structure Incorporating Microstructure and Surface <span class="hlt">Energy</span> Effects</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Zhang, G. Y.; Gao, X. -L.; Bishop, J. E.; ...</p> <p>2017-11-20</p> <p>Here, a new model for determining <span class="hlt">band</span> gaps for elastic wave propagation in a periodic composite beam structure is developed using a non-classical Bernoulli–Euler beam model that incorporates the microstructure, surface <span class="hlt">energy</span> and rotational inertia effects. The Bloch theorem and transfer matrix method for periodic structures are employed in the formulation. The new model reduces to the classical elasticity-based model when both the microstructure and surface <span class="hlt">energy</span> effects are not considered. The <span class="hlt">band</span> gaps predicted by the new model depend on the microstructure and surface elasticity of each constituent material, the unit cell size, the rotational inertia, and the volumemore » fraction. To quantitatively illustrate the effects of these factors, a parametric study is conducted. The numerical results reveal that the <span class="hlt">band</span> gap predicted by the current non-classical model is always larger than that predicted by the classical model when the beam thickness is very small, but the difference is diminishing as the thickness becomes large. Also, it is found that the first frequency for producing the <span class="hlt">band</span> gap and the <span class="hlt">band</span> gap size decrease with the increase of the unit cell length according to both the current and classical models. In <span class="hlt">addition</span>, it is observed that the effect of the rotational inertia is larger when the exciting frequency is higher and the unit cell length is smaller. Furthermore, it is seen that the volume fraction has a significant effect on the <span class="hlt">band</span> gap size, and large <span class="hlt">band</span> gaps can be obtained by tailoring the volume fraction and material parameters.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011JAP...109k3724M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011JAP...109k3724M"><span>Branch-point <span class="hlt">energies</span> and the <span class="hlt">band</span>-structure lineup at Schottky contacts and heterostrucures</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mönch, Winfried</p> <p>2011-06-01</p> <p>Empirical branch-point <span class="hlt">energies</span> of Si, the group-III nitrides AlN, GaN, and InN, and the group-II and group-III oxides MgO, ZnO, Al2O3 and In2O3 are determined from experimental valance-<span class="hlt">band</span> offsets of their heterostructures. For Si, GaN, and MgO, these values agree with the branch-point <span class="hlt">energies</span> obtained from the barrier heights of their Schottky contacts. The empirical branch-point <span class="hlt">energies</span> of Si and the group-III nitrides are in very good agreement with results of previously published calculations using quite different approaches such as the empirical tight-binding approximation and modern electronic-structure theory. In contrast, the empirical branch-point <span class="hlt">energies</span> of the group-II and group-III oxides do not confirm the respective theoretical results. As at Schottky contacts, the <span class="hlt">band</span>-structure lineup at heterostructures is also made up of a zero-charge-transfer term and an intrinsic electric-dipole contribution. Hence, valence-<span class="hlt">band</span> offsets are not equal to the difference of the branch-point <span class="hlt">energies</span> of the two semiconductors forming the heterostructure. The electric-dipole term may be described by the electronegativity difference of the two solids in contact. A detailed analysis of experimental Si Schottky barrier heights and heterostructure valence-<span class="hlt">band</span> offsets explains and proves these conclusions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015JMPSo..75...45X','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015JMPSo..75...45X"><span>Atomistic potentials based <span class="hlt">energy</span> flux integral criterion for dynamic adiabatic shear <span class="hlt">banding</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Xu, Yun; Chen, Jun</p> <p>2015-02-01</p> <p>The <span class="hlt">energy</span> flux integral criterion based on atomistic potentials within the framework of hyperelasticity-plasticity is proposed for dynamic adiabatic shear <span class="hlt">banding</span> (ASB). System Helmholtz <span class="hlt">energy</span> decomposition reveals that the dynamic influence on the integral path dependence is originated from the volumetric strain <span class="hlt">energy</span> and partial deviatoric strain <span class="hlt">energy</span>, and the plastic influence only from the rest part of deviatoric strain <span class="hlt">energy</span>. The concept of critical shear <span class="hlt">banding</span> <span class="hlt">energy</span> is suggested for describing the initiation of ASB, which consists of the dynamic recrystallization (DRX) threshold <span class="hlt">energy</span> and the thermal softening <span class="hlt">energy</span>. The criterion directly relates <span class="hlt">energy</span> flux to the basic physical processes that induce shear instability such as dislocation nucleations and multiplications, without introducing ad-hoc parameters in empirical constitutive models. It reduces to the classical path independent J-integral for quasi-static loading and elastic solids. The atomistic-to-continuum multiscale coupling method is used to simulate the initiation of ASB. Atomic configurations indicate that DRX induced microstructural softening may be essential to the dynamic shear localization and hence the initiation of ASB.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018AIPC.1935f0001A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AIPC.1935f0001A"><span>Design of a dual <span class="hlt">band</span> metamaterial absorber for Wi-Fi <span class="hlt">bands</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Alkurt, Fatih Özkan; Baǧmancı, Mehmet; Karaaslan, Muharrem; Bakır, Mehmet; Altıntaş, Olcay; Karadaǧ, Faruk; Akgöl, Oǧuzhan; Ünal, Emin</p> <p>2018-02-01</p> <p>The goal of this work is to design and fabrication of a dual <span class="hlt">band</span> metamaterial based absorber for Wireless Fidelity (Wi-Fi) <span class="hlt">bands</span>. Wi-Fi has two different operating frequencies such as 2.45 GHz and 5 GHz. A dual <span class="hlt">band</span> absorber is proposed and the proposed structure consists of two layered unit cells, and different sized square split ring (SSR) resonators located on each layers. Copper is used for metal layer and resonator structure, FR-4 is used as substrate layer in the proposed structure. This designed dual <span class="hlt">band</span> metamaterial absorber is used in the wireless frequency <span class="hlt">bands</span> which has two center frequencies such as 2.45 GHz and 5 GHz. Finite Integration Technique (FIT) based simulation software used and according to FIT based simulation results, the absorption peak in the 2.45 GHz is about 90% and the another frequency 5 GHz has absorption peak near 99%. In <span class="hlt">addition</span>, this proposed structure has a potential for <span class="hlt">energy</span> harvesting applications in future works.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28402113','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28402113"><span>Importance of the Kinetic <span class="hlt">Energy</span> Density for <span class="hlt">Band</span> Gap Calculations in Solids with Density Functional Theory.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Tran, Fabien; Blaha, Peter</p> <p>2017-05-04</p> <p>Recently, exchange-correlation potentials in density functional theory were developed with the goal of providing improved <span class="hlt">band</span> gaps in solids. Among them, the semilocal potentials are particularly interesting for large systems since they lead to calculations that are much faster than with hybrid functionals or methods like GW. We present an exhaustive comparison of semilocal exchange-correlation potentials for <span class="hlt">band</span> gap calculations on a large test set of solids, and particular attention is paid to the potential HLE16 proposed by Verma and Truhlar. It is shown that the most accurate potential is the modified Becke-Johnson potential, which, most noticeably, is much more accurate than all other semilocal potentials for strongly correlated systems. This can be attributed to its <span class="hlt">additional</span> dependence on the kinetic <span class="hlt">energy</span> density. It is also shown that the modified Becke-Johnson potential is at least as accurate as the hybrid functionals and more reliable for solids with large <span class="hlt">band</span> gaps.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2007EPJB...59..391D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2007EPJB...59..391D"><span><span class="hlt">Energy</span> diffusion controlled reaction rate of reacting particle driven by broad-<span class="hlt">band</span> noise</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Deng, M. L.; Zhu, W. Q.</p> <p>2007-10-01</p> <p>The <span class="hlt">energy</span> diffusion controlled reaction rate of a reacting particle with linear weak damping and broad-<span class="hlt">band</span> noise excitation is studied by using the stochastic averaging method. First, the stochastic averaging method for strongly nonlinear oscillators under broad-<span class="hlt">band</span> noise excitation using generalized harmonic functions is briefly introduced. Then, the reaction rate of the classical Kramers' reacting model with linear weak damping and broad-<span class="hlt">band</span> noise excitation is investigated by using the stochastic averaging method. The averaged Itô stochastic differential equation describing the <span class="hlt">energy</span> diffusion and the Pontryagin equation governing the mean first-passage time (MFPT) are established. The <span class="hlt">energy</span> diffusion controlled reaction rate is obtained as the inverse of the MFPT by solving the Pontryagin equation. The results of two special cases of broad-<span class="hlt">band</span> noises, i.e. the harmonic noise and the exponentially corrected noise, are discussed in details. It is demonstrated that the general expression of reaction rate derived by the authors can be reduced to the classical ones via linear approximation and high potential barrier approximation. The good agreement with the results of the Monte Carlo simulation verifies that the reaction rate can be well predicted using the stochastic averaging method.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018AIPC.1953i0066S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AIPC.1953i0066S"><span>Optical absorption spectra and <span class="hlt">energy</span> <span class="hlt">band</span> gap in manganese containing sodium zinc phosphate glasses</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sardarpasha, K. R.; Hanumantharaju, N.; Gowda, V. C. Veeranna</p> <p>2018-05-01</p> <p>Optical <span class="hlt">band</span> gap <span class="hlt">energy</span> in the system 25Na2O-(75-x)[0.6P2O5-0.4ZnO]-xMnO2 (where x = 0.5,1,5,10 and 20 mol.%) have been studied. The intensity of the absorption <span class="hlt">band</span> found to increase with increase of MnO2 content. The decrease in the optical <span class="hlt">band</span> gap <span class="hlt">energy</span> with increase in MnO2 content in the investigated glasses is attributed to shifting of absorption edge to a longer wavelength region. The obtained results were discussed in view of the structure of phosphate glass network.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017Nanot..28s5604O','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017Nanot..28s5604O"><span>Tuning Ferritin’s <span class="hlt">band</span> gap through mixed metal oxide nanoparticle formation</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Olsen, Cameron R.; Embley, Jacob S.; Hansen, Kameron R.; Henrichsen, Andrew M.; Peterson, J. Ryan; Colton, John S.; Watt, Richard K.</p> <p>2017-05-01</p> <p>This study uses the formation of a mixed metal oxide inside ferritin to tune the <span class="hlt">band</span> gap <span class="hlt">energy</span> of the ferritin mineral. The mixed metal oxide is composed of both Co and Mn, and is formed by reacting aqueous Co2+ with {{{{MnO}}}4}- in the presence of apoferritin. Altering the ratio between the two reactants allowed for controlled tuning of the <span class="hlt">band</span> gap <span class="hlt">energies</span>. All minerals formed were indirect <span class="hlt">band</span> gap materials, with indirect <span class="hlt">band</span> gap <span class="hlt">energies</span> ranging from 0.52 to 1.30 eV. The direct transitions were also measured, with <span class="hlt">energy</span> values ranging from 2.71 to 3.11 eV. Tuning the <span class="hlt">band</span> gap <span class="hlt">energies</span> of these samples changes the wavelengths absorbed by each mineral, increasing ferritin’s potential in solar-<span class="hlt">energy</span> harvesting. <span class="hlt">Additionally</span>, the success of using {{{{MnO}}}4}- in ferritin mineral formation opens the possibility for new mixed metal oxide cores inside ferritin.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1988AcSpA..44..505S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1988AcSpA..44..505S"><span>Potential <span class="hlt">energy</span> surface and vibrational <span class="hlt">band</span> origins of the triatomic lithium cation</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Searles, Debra J.; Dunne, Simon J.; von Nagy-Felsobuki, Ellak I.</p> <p></p> <p>The 104 point CISD Li +3 potential <span class="hlt">energy</span> surface and its analytical representation is reported. The calculations predict the minimum <span class="hlt">energy</span> geometry to be an equilateral triangle of side RLiLi = 3.0 Å and of <span class="hlt">energy</span> - 22.20506 E h. A fifth-order Morse—Dunham type analytical force field is used in the Carney—Porter normal co-ordinate vibrational Hamiltonian, the corresponding eigenvalue problem being solved variationally using a 560 configurational finite-element basis set. The predicted assignment of the vibrational <span class="hlt">band</span> origins is in accord with that reported for H +3. Moreover, for 6Li +3 and 7Li +3 the lowest i.r. accessible <span class="hlt">band</span> origin is the overlineν0,1,±1 predicted to be at 243.6 and 226.0 cm -1 respectively.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018AIPC.1942e0111S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AIPC.1942e0111S"><span>Determination of shift in <span class="hlt">energy</span> of <span class="hlt">band</span> edges and <span class="hlt">band</span> gap of ZnSe spherical quantum dot</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Siboh, Dutem; Kalita, Pradip Kumar; Sarma, Jayanta Kumar; Nath, Nayan Mani</p> <p>2018-04-01</p> <p>We have determined the quantum confinement induced shifts in <span class="hlt">energy</span> of <span class="hlt">band</span> edges and <span class="hlt">band</span> gap with respect to size of ZnSe spherical quantum dot employing an effective confinement potential model developed in our earlier communication "arXiv:1705.10343". We have also performed phenomenological analysis of our theoretical results in comparison with available experimental data and observe a very good agreement in this regard. Phenomenological success achieved in this regard confirms validity of the confining potential model as well as signifies the capability and applicability of the ansatz for the effective confining potential to have reasonable information in the study of real nano-structured spherical systems.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014JAP...115n3107D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014JAP...115n3107D"><span>Inter-<span class="hlt">band</span> optoelectronic properties in quantum dot structure of low <span class="hlt">band</span> gap III-V semiconductors</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Dey, Anup; Maiti, Biswajit; Chanda Sarkar, Debasree</p> <p>2014-04-01</p> <p>A generalized theory is developed to study inter-<span class="hlt">band</span> optical absorption coefficient (IOAC) and material gain (MG) in quantum dot structures of narrow gap III-V compound semiconductor considering the wave-vector (k→) dependence of the optical transition matrix element. The <span class="hlt">band</span> structures of these low <span class="hlt">band</span> gap semiconducting materials with sufficiently separated split-off valance <span class="hlt">band</span> are frequently described by the three <span class="hlt">energy</span> <span class="hlt">band</span> model of Kane. This has been adopted for analysis of the IOAC and MG taking InAs, InSb, Hg1-xCdxTe, and In1-xGaxAsyP1-y lattice matched to InP, as example of III-V compound semiconductors, having varied split-off <span class="hlt">energy</span> <span class="hlt">band</span> compared to their bulk <span class="hlt">band</span> gap <span class="hlt">energy</span>. It has been found that magnitude of the IOAC for quantum dots increases with increasing incident photon <span class="hlt">energy</span> and the lines of absorption are more closely spaced in the three <span class="hlt">band</span> model of Kane than those with parabolic <span class="hlt">energy</span> <span class="hlt">band</span> approximations reflecting the direct the influence of <span class="hlt">energy</span> <span class="hlt">band</span> parameters. The results show a significant deviation to the MG spectrum of narrow-gap materials having <span class="hlt">band</span> nonparabolicity compared to the parabolic <span class="hlt">band</span> model approximations. The results reflect the important role of valence <span class="hlt">band</span> split-off <span class="hlt">energies</span> in these narrow gap semiconductors.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JAP...123n5111G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JAP...123n5111G"><span>The temperature-dependency of the optical <span class="hlt">band</span> gap of ZnO measured by electron <span class="hlt">energy</span>-loss spectroscopy in a scanning transmission electron microscope</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Granerød, Cecilie S.; Galeckas, Augustinas; Johansen, Klaus Magnus; Vines, Lasse; Prytz, Øystein</p> <p>2018-04-01</p> <p>The optical <span class="hlt">band</span> gap of ZnO has been measured as a function of temperature using Electron <span class="hlt">Energy</span>-Loss Spectroscopy (EELS) in a (Scanning) Transmission Electron Microscope ((S)TEM) from approximately 100 K up towards 1000 K. The <span class="hlt">band</span> gap narrowing shows a close to linear dependency for temperatures above 250 K and is accurately described by Varshni, Bose-Einstein, Pässler and Manoogian-Woolley models. <span class="hlt">Additionally</span>, the measured <span class="hlt">band</span> gap is compared with both optical absorption measurements and photoluminescence data. STEM-EELS is here shown to be a viable technique to measure optical <span class="hlt">band</span> gaps at elevated temperatures, with an available temperature range up to 1500 K and the benefit of superior spatial resolution.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26247853','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26247853"><span><span class="hlt">Energy</span> Impacts of Wide <span class="hlt">Band</span> Gap Semiconductors in U.S. Light-Duty Electric Vehicle Fleet.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Warren, Joshua A; Riddle, Matthew E; Graziano, Diane J; Das, Sujit; Upadhyayula, Venkata K K; Masanet, Eric; Cresko, Joe</p> <p>2015-09-01</p> <p>Silicon carbide and gallium nitride, two leading wide <span class="hlt">band</span> gap semiconductors with significant potential in electric vehicle power electronics, are examined from a life cycle <span class="hlt">energy</span> perspective and compared with incumbent silicon in U.S. light-duty electric vehicle fleet. Cradle-to-gate, silicon carbide is estimated to require more than twice the <span class="hlt">energy</span> as silicon. However, the magnitude of vehicle use phase fuel savings potential is comparatively several orders of magnitude higher than the marginal increase in cradle-to-gate <span class="hlt">energy</span>. Gallium nitride cradle-to-gate <span class="hlt">energy</span> requirements are estimated to be similar to silicon, with use phase savings potential similar to or exceeding that of silicon carbide. Potential <span class="hlt">energy</span> reductions in the United States vehicle fleet are examined through several scenarios that consider the market adoption potential of electric vehicles themselves, as well as the market adoption potential of wide <span class="hlt">band</span> gap semiconductors in electric vehicles. For the 2015-2050 time frame, cumulative <span class="hlt">energy</span> savings associated with the deployment of wide <span class="hlt">band</span> gap semiconductors are estimated to range from 2-20 billion GJ depending on market adoption dynamics.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_1");'>1</a></li> <li class="active"><span>2</span></li> <li><a href="#" onclick='return showDiv("page_3");'>3</a></li> <li><a href="#" onclick='return showDiv("page_4");'>4</a></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_2 --> <div id="page_3" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_1");'>1</a></li> <li><a href="#" onclick='return showDiv("page_2");'>2</a></li> <li class="active"><span>3</span></li> <li><a href="#" onclick='return showDiv("page_4");'>4</a></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="41"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017PhDT.......102T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017PhDT.......102T"><span>Electronic <span class="hlt">Band</span> Structure Tuning of Highly-Mismatched-Alloys for <span class="hlt">Energy</span> Conversion Applications</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ting, Min</p> <p></p> <p>Highly-mismatched alloys: ZnO1-xTe x and GaN1-xSb x are discussed within the context of finding the suitable material for a cost-effective Si-based tandem solar cell (SBTSC). SBTSC is an attractive concept for breaking through the <span class="hlt">energy</span> conversion efficiency theoretical limit of a single junction solar cell. Combining with a material of 1.8 eV <span class="hlt">band</span> gap, SBTSC can theoretically achieve <span class="hlt">energy</span> conversion efficiency > 45%. ZnO and GaN are wide <span class="hlt">band</span> gap semiconductors. Alloying Te in ZnO and alloying Sb in GaN result in large <span class="hlt">band</span> gap reduction to < 2 eV from 3.3 eV and 3.4 eV respectively. The <span class="hlt">band</span> gap reduction is majorly achieved by the upward shift of valence <span class="hlt">band</span> (VB). Incorporating Te in ZnO modifies the VB of ZnO through the valence-<span class="hlt">band</span> anticrossing (VBAC) interaction between localized Te states and ZnO VB delocalized states, which forms a Te-derived VB at 1 eV above the host VB. Similar <span class="hlt">band</span> structure modification is resulted from alloying Sb in GaN. Zn1-xTex and GaN 1-xSbx thin films are synthesized across the whole composition range by pulsed laser deposition (PLD) and low temperature molecular beam epitaxy (LT-MBE) respectively. The electronic <span class="hlt">band</span> edges of these alloys are measured by synchrotron X-ray absorption, emission, and the X-ray photoelectron spectroscopies. Modeling the optical absorption coefficient with the <span class="hlt">band</span> anticrossing (BAC) model revealed that the Te and Sb defect levels to be at 0.99 eV and 1.2 eV above the VB of ZnO and GaN respectively. Electrically, Zn1-xTex is readily n-type conductive and GaN1-xSbx is strongly p-type conductive. A heterojunction device of p-type GaN 0.93Sb0.07 with n-type ZnO0.77Te0.93 upper cell (<span class="hlt">band</span> gap at 1.8 eV) on Si bottom cell is proposed as a promising SBTSC device.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016JPhD...49b5502W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016JPhD...49b5502W"><span>Generalized thermoelastic wave <span class="hlt">band</span> gaps in phononic crystals without <span class="hlt">energy</span> dissipation</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wu, Ying; Yu, Kaiping; Li, Xiao; Zhou, Haotian</p> <p>2016-01-01</p> <p>We present a theoretical investigation of the thermoelastic wave propagation in the phononic crystals in the context of Green-Nagdhi theory by taking thermoelastic coupling into account. The thermal field is assumed to be steady. Thermoelastic wave <span class="hlt">band</span> structures of 3D and 2D are derived by using the plane wave expansion method. For the 2D problem, the anti-plane shear mode is not affected by the temperature difference. Thermoelastic wave <span class="hlt">bands</span> of the in-plane x-y mode are calculated for lead/silicone rubber, aluminium/silicone rubber, and aurum/silicone rubber phononic crystals. The new findings in the numerical results indicate that the thermoelastic wave <span class="hlt">bands</span> are composed of the pure elastic wave <span class="hlt">bands</span> and the thermal wave <span class="hlt">bands</span>, and that the thermal wave <span class="hlt">bands</span> can serve as the low boundary of the first <span class="hlt">band</span> gap when the filling ratio is low. In <span class="hlt">addition</span>, for the lead/silicone rubber phononic crystals the effects of lattice type (square, rectangle, regular triangle, and hexagon) and inclusion shape (circle, oval, and square) on the normalized thermoelastic bandwidth and the upper/lower gap boundaries are analysed and discussed. It is concluded that their effects on the thermoelastic wave <span class="hlt">band</span> structure are remarkable.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2004PhDT........23T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2004PhDT........23T"><span><span class="hlt">Energies</span> of rare-earth ion states relative to host <span class="hlt">bands</span> in optical materials from electron photoemission spectroscopy</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Thiel, Charles Warren</p> <p></p> <p>There are a vast number of applications for rare-earth-activated materials and much of today's cutting-edge optical technology and emerging innovations are enabled by their unique properties. In many of these applications, interactions between the rare-earth ion and the host material's electronic states can enhance or inhibit performance and provide mechanisms for manipulating the optical properties. Continued advances in these technologies require knowledge of the relative <span class="hlt">energies</span> of rare-earth and crystal <span class="hlt">band</span> states so that properties of available materials may be fully understood and new materials may be logically developed. Conventional and resonant electron photoemission techniques were used to measure 4f electron and valence <span class="hlt">band</span> binding <span class="hlt">energies</span> in important optical materials, including YAG, YAlO3, and LiYF4. The photoemission spectra were theoretically modeled and analyzed to accurately determine relative <span class="hlt">energies</span>. By combining these <span class="hlt">energies</span> with ultraviolet spectroscopy, binding <span class="hlt">energies</span> of excited 4fN-15d and 4fN+1 states were determined. While the 4fN ground-state <span class="hlt">energies</span> vary considerably between different trivalent ions and lie near or below the top of the valence <span class="hlt">band</span> in optical materials, the lowest 4f N-15d states have similar <span class="hlt">energies</span> and are near the bottom of the conduction <span class="hlt">band</span>. As an example for YAG, the Tb3+ 4f N ground state is in the <span class="hlt">band</span> gap at 0.7 eV above the valence <span class="hlt">band</span> while the Lu3+ ground state is 4.7 eV below the valence <span class="hlt">band</span> maximum; however, the lowest 4fN-15d states are 2.2 eV below the conduction <span class="hlt">band</span> for both ions. We found that a simple model accurately describes the binding <span class="hlt">energies</span> of the 4fN, 4fN-1 5d, and 4fN+1 states. The model's success across the entire rare-earth series indicates that measurements on two different ions in a host are sufficient to predict the <span class="hlt">energies</span> of all rare-earth ions in that host. This information provides new insight into electron transfer transitions, luminescence quenching, and valence</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011AGUFMED41A0490H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011AGUFMED41A0490H"><span>Examining the Displacement of <span class="hlt">Energy</span> during Formation of Shear <span class="hlt">Bands</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hernandez, M.; Hilley, G. E.</p> <p>2011-12-01</p> <p>M.X. Hernandez, G. Hilley Department of Geological and Environmental Sciences, Stanford University, Stanford, CA This study has originated from an experimental (sandbox) setting that we have previously used to document the link between the kinematics and dynamics of deforming sand in the verge of frictional failure. Our initial experimental setting included a load control system that allowed us to track the changes in load, that when applied to the sand, deform and generate individual shear <span class="hlt">bands</span> or localized faults. Over the course of earlier experiments, three cameras located at different positions outside the sandbox monitored the movement throughout the run. This current stage of analysis includes using computer programs such as QuickTime to create image sequences of the shear <span class="hlt">band</span> formation, and Microsoft Excel to visually graph and plot each data sequence. This allows us to investigate the correlation between changes in work measured within our experiments, the construction of topography, slip along shear <span class="hlt">bands</span>, and the creation of new shear <span class="hlt">bands</span>. We observed that the measured load generally increased during the experiment to maintain a constant displacement rate as the sand wedge thickened and modeled topography increased. Superposed on this trend were periodic drops in load that appeared temporally coincident with the formation of shear <span class="hlt">bands</span> in the sand. Using the time series of the loads applied during the experiment, changes in the position of the backstop over time, and the loads measured before, during, and after the time of each shear <span class="hlt">band</span> formation, we are examining the fraction of the apples work that is absorbed by friction and shear <span class="hlt">band</span> formation, and what fraction of the apples work is expended in increasing the potential <span class="hlt">energy</span> of the thickening sand wedge. Our results indicate that before the formation of a continuous shear <span class="hlt">band</span>, the rate of work done on the sand by the experimental apparatus decreases. This may suggest that once formed, work</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/1414278-energy-impacts-wide-band-gap-semiconductors-light-duty-electric-vehicle-fleet','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/1414278-energy-impacts-wide-band-gap-semiconductors-light-duty-electric-vehicle-fleet"><span><span class="hlt">Energy</span> Impacts of Wide <span class="hlt">Band</span> Gap Semiconductors in U.S. Light-Duty Electric Vehicle Fleet</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Warren, Joshua A.; Riddle, Matthew E.; Graziano, Diane J.</p> <p>2015-08-12</p> <p>Silicon carbide and gallium nitride, two leading wide <span class="hlt">band</span> gap semiconductors with significant potential in electric vehicle power electronics, are examined from a life cycle <span class="hlt">energy</span> perspective and compared with incumbent silicon in U.S. light-duty electric vehicle fleet. Cradle-to-gate, silicon carbide is estimated to require more than twice the <span class="hlt">energy</span> as silicon. However, the magnitude of vehicle use phase fuel savings potential is comparatively several orders of magnitude higher than the marginal increase in cradle-to-gate <span class="hlt">energy</span>. Gallium nitride cradle-to-gate <span class="hlt">energy</span> requirements are estimated to be similar to silicon, with use phase savings potential similar to or exceeding that of siliconmore » carbide. Potential <span class="hlt">energy</span> reductions in the United States vehicle fleet are examined through several scenarios that consider the market adoption potential of electric vehicles themselves, as well as the market adoption potential of wide <span class="hlt">band</span> gap semiconductors in electric vehicles. For the 2015–2050 time frame, cumulative <span class="hlt">energy</span> savings associated with the deployment of wide <span class="hlt">band</span> gap semiconductors are estimated to range from 2–20 billion GJ depending on market adoption dynamics.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015PhRvB..92u4514K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015PhRvB..92u4514K"><span>Fragile surface zero-<span class="hlt">energy</span> flat <span class="hlt">bands</span> in three-dimensional chiral superconductors</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kobayashi, Shingo; Tanaka, Yukio; Sato, Masatoshi</p> <p>2015-12-01</p> <p>We study surface zero-<span class="hlt">energy</span> flat <span class="hlt">bands</span> in three-dimensional chiral superconductors with pz(px+i py) ν -wave pairing symmetry (ν is a nonzero integer), based on topological arguments and tunneling conductance. It is shown that the surface flat <span class="hlt">bands</span> are fragile against (i) the surface misorientation and (ii) the surface Rashba spin-orbit interaction. The fragility of (i) is specific to chiral SCs, whereas that of (ii) happens for general odd-parity SCs. We demonstrate that these flat-<span class="hlt">band</span> instabilities vanish or suppress a zero-bias conductance peak in a normal/insulator/superconductor junction, which behavior is clearly different from high-Tc cuprates and noncentrosymmetric superconductors. By calculating the angle-resolved conductance, we also discuss a topological surface state associated with the coexistence of line and point nodes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23214551','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23214551"><span><span class="hlt">Energy</span> transport in weakly nonlinear wave systems with narrow frequency <span class="hlt">band</span> excitation.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kartashova, Elena</p> <p>2012-10-01</p> <p>A novel discrete model (D model) is presented describing nonlinear wave interactions in systems with small and moderate nonlinearity under narrow frequency <span class="hlt">band</span> excitation. It integrates in a single theoretical frame two mechanisms of <span class="hlt">energy</span> transport between modes, namely, intermittency and <span class="hlt">energy</span> cascade, and gives the conditions under which each regime will take place. Conditions for the formation of a cascade, cascade direction, conditions for cascade termination, etc., are given and depend strongly on the choice of excitation parameters. The <span class="hlt">energy</span> spectra of a cascade may be computed, yielding discrete and continuous <span class="hlt">energy</span> spectra. The model does not require statistical assumptions, as all effects are derived from the interaction of distinct modes. In the example given-surface water waves with dispersion function ω(2)=gk and small nonlinearity-the D model predicts asymmetrical growth of side-<span class="hlt">bands</span> for Benjamin-Feir instability, while the transition from discrete to continuous <span class="hlt">energy</span> spectrum, excitation parameters properly chosen, yields the saturated Phillips' power spectrum ~g(2)ω(-5). The D model can be applied to the experimental and theoretical study of numerous wave systems appearing in hydrodynamics, nonlinear optics, electrodynamics, plasma, convection theory, etc.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27608986','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27608986"><span>Free-end adaptive nudged elastic <span class="hlt">band</span> method for locating transition states in minimum <span class="hlt">energy</span> path calculation.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zhang, Jiayong; Zhang, Hongwu; Ye, Hongfei; Zheng, Yonggang</p> <p>2016-09-07</p> <p>A free-end adaptive nudged elastic <span class="hlt">band</span> (FEA-NEB) method is presented for finding transition states on minimum <span class="hlt">energy</span> paths, where the <span class="hlt">energy</span> barrier is very narrow compared to the whole paths. The previously proposed free-end nudged elastic <span class="hlt">band</span> method may suffer from convergence problems because of the kinks arising on the elastic <span class="hlt">band</span> if the initial elastic <span class="hlt">band</span> is far from the minimum <span class="hlt">energy</span> path and weak springs are adopted. We analyze the origin of the formation of kinks and present an improved free-end algorithm to avoid the convergence problem. Moreover, by coupling the improved free-end algorithm and an adaptive strategy, we develop a FEA-NEB method to accurately locate the transition state with the elastic <span class="hlt">band</span> cut off repeatedly and the density of images near the transition state increased. Several representative numerical examples, including the dislocation nucleation in a penta-twinned nanowire, the twin boundary migration under a shear stress, and the cross-slip of screw dislocation in face-centered cubic metals, are investigated by using the FEA-NEB method. Numerical results demonstrate both the stability and efficiency of the proposed method.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1256092','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1256092"><span><span class="hlt">Additive</span> Manufacturing Integrated <span class="hlt">Energy</span> Demonstration</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Jackson, Roderick; Lee, Brian; Love, Lonnie</p> <p>2016-02-05</p> <p>Meet AMIE - the <span class="hlt">Additive</span> Manufacturing Integrated <span class="hlt">Energy</span> demonstration project. Led by Oak Ridge National Laboratory and many industry partners, the AMIE project changes the way we think about generating, storing, and using electrical power. AMIE uses an integrated <span class="hlt">energy</span> system that shares <span class="hlt">energy</span> between a building and a vehicle. And, utilizing advanced manufacturing and rapid innovation, it only took one year from concept to launch.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.osti.gov/sciencecinema/biblio/1256092','SCIGOVIMAGE-SCICINEMA'); return false;" href="http://www.osti.gov/sciencecinema/biblio/1256092"><span><span class="hlt">Additive</span> Manufacturing Integrated <span class="hlt">Energy</span> Demonstration</span></a></p> <p><a target="_blank" href="http://www.osti.gov/sciencecinema/">ScienceCinema</a></p> <p>Jackson, Roderick; Lee, Brian; Love, Lonnie; Mabe, Gavin; Keller, Martin; Curran, Scott; Chinthavali, Madhu; Green, Johney; Sawyer, Karma; Enquist, Phil</p> <p>2018-01-16</p> <p>Meet AMIE - the <span class="hlt">Additive</span> Manufacturing Integrated <span class="hlt">Energy</span> demonstration project. Led by Oak Ridge National Laboratory and many industry partners, the AMIE project changes the way we think about generating, storing, and using electrical power. AMIE uses an integrated <span class="hlt">energy</span> system that shares <span class="hlt">energy</span> between a building and a vehicle. And, utilizing advanced manufacturing and rapid innovation, it only took one year from concept to launch.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5644240','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5644240"><span><span class="hlt">Additive</span> Manufacturing: Unlocking the Evolution of <span class="hlt">Energy</span> Materials</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Zhakeyev, Adilet; Wang, Panfeng; Shu, Wenmiao; Wang, Huizhi</p> <p>2017-01-01</p> <p>Abstract The global <span class="hlt">energy</span> infrastructure is undergoing a drastic transformation towards renewable <span class="hlt">energy</span>, posing huge challenges on the <span class="hlt">energy</span> materials research, development and manufacturing. <span class="hlt">Additive</span> manufacturing has shown its promise to change the way how future <span class="hlt">energy</span> system can be designed and delivered. It offers capability in manufacturing complex 3D structures, with near‐complete design freedom and high sustainability due to minimal use of materials and toxic chemicals. Recent literatures have reported that <span class="hlt">additive</span> manufacturing could unlock the evolution of <span class="hlt">energy</span> materials and chemistries with unprecedented performance in the way that could never be achieved by conventional manufacturing techniques. This comprehensive review will fill the gap in communicating on recent breakthroughs in <span class="hlt">additive</span> manufacturing for <span class="hlt">energy</span> material and device applications. It will underpin the discoveries on what 3D functional <span class="hlt">energy</span> structures can be created without design constraints, which bespoke <span class="hlt">energy</span> materials could be <span class="hlt">additively</span> manufactured with customised solutions, and how the <span class="hlt">additively</span> manufactured devices could be integrated into <span class="hlt">energy</span> systems. This review will also highlight emerging and important applications in <span class="hlt">energy</span> <span class="hlt">additive</span> manufacturing, including fuel cells, batteries, hydrogen, solar cell as well as carbon capture and storage. PMID:29051861</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29051861','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29051861"><span><span class="hlt">Additive</span> Manufacturing: Unlocking the Evolution of <span class="hlt">Energy</span> Materials.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zhakeyev, Adilet; Wang, Panfeng; Zhang, Li; Shu, Wenmiao; Wang, Huizhi; Xuan, Jin</p> <p>2017-10-01</p> <p>The global <span class="hlt">energy</span> infrastructure is undergoing a drastic transformation towards renewable <span class="hlt">energy</span>, posing huge challenges on the <span class="hlt">energy</span> materials research, development and manufacturing. <span class="hlt">Additive</span> manufacturing has shown its promise to change the way how future <span class="hlt">energy</span> system can be designed and delivered. It offers capability in manufacturing complex 3D structures, with near-complete design freedom and high sustainability due to minimal use of materials and toxic chemicals. Recent literatures have reported that <span class="hlt">additive</span> manufacturing could unlock the evolution of <span class="hlt">energy</span> materials and chemistries with unprecedented performance in the way that could never be achieved by conventional manufacturing techniques. This comprehensive review will fill the gap in communicating on recent breakthroughs in <span class="hlt">additive</span> manufacturing for <span class="hlt">energy</span> material and device applications. It will underpin the discoveries on what 3D functional <span class="hlt">energy</span> structures can be created without design constraints, which bespoke <span class="hlt">energy</span> materials could be <span class="hlt">additively</span> manufactured with customised solutions, and how the <span class="hlt">additively</span> manufactured devices could be integrated into <span class="hlt">energy</span> systems. This review will also highlight emerging and important applications in <span class="hlt">energy</span> <span class="hlt">additive</span> manufacturing, including fuel cells, batteries, hydrogen, solar cell as well as carbon capture and storage.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22421987','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22421987"><span>Nanoscale charge distribution and <span class="hlt">energy</span> <span class="hlt">band</span> modification in defect-patterned graphene.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wang, Shengnan; Wang, Rui; Wang, Xiaowei; Zhang, Dongdong; Qiu, Xiaohui</p> <p>2012-04-21</p> <p>Defects were introduced precisely to exfoliated graphene (G) sheets on a SiO(2)/n(+) Si substrate to modulate the local <span class="hlt">energy</span> <span class="hlt">band</span> structure and the electron pathway using solution-phase oxidation followed by thermal reduction. The resulting nanoscale charge distribution and <span class="hlt">band</span> gap modification were investigated by electrostatic force microscopy and spectroscopy. A transition phase with coexisting submicron-sized metallic and insulating regions in the moderately oxidized monolayer graphene were visualized and measured directly. It was determined that the delocalization of electrons/holes in a graphene "island" is confined by the surrounding defective C-O matrix, which acts as an <span class="hlt">energy</span> barrier for mobile charge carriers. In contrast to the irreversible structural variations caused by the oxidation process, the electrical properties of graphene can be restored by annealing. The defect-patterned graphene and graphene oxide heterojunctions were further characterized by electrical transport measurement.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1176927','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1176927"><span>Strategic <span class="hlt">Energy</span> Management Plan for the Santa Ynez <span class="hlt">Band</span> of Chumash Indians</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Davenport, Lars; Smythe, Louisa; Sarquilla, Lindsey</p> <p>2015-03-27</p> <p>This plan outlines the Santa Ynez <span class="hlt">Band</span> of Chumash Indians’ comprehensive <span class="hlt">energy</span> management strategy including an assessment of current practices, a commitment to improving <span class="hlt">energy</span> performance and reducing overall <span class="hlt">energy</span> use, and recommended actions to achieve these goals. Vision Statement The primary objective of the Strategic <span class="hlt">Energy</span> Management Plan is to implement <span class="hlt">energy</span> efficiency, <span class="hlt">energy</span> security, conservation, education, and renewable <span class="hlt">energy</span> projects that align with the economic goals and cultural values of the community to improve the health and welfare of the tribe. The intended outcomes of implementing the <span class="hlt">energy</span> plan include job creation, capacity building, and reduced <span class="hlt">energy</span> costsmore » for tribal community members, and tribal operations. By encouraging <span class="hlt">energy</span> independence and local power production the plan will promote self-sufficiency. Mission & Objectives The Strategic <span class="hlt">Energy</span> Plan will provide information and suggestions to guide tribal decision-making and provide a foundation for effective management of <span class="hlt">energy</span> resources within the Santa Ynez <span class="hlt">Band</span> of Chumash Indians (SYBCI) community. The objectives of developing this plan include; Assess current <span class="hlt">energy</span> demand and costs of all tribal enterprises, offices, and facilities; Provide a baseline assessment of the SYBCI’s <span class="hlt">energy</span> resources so that future progress can be clearly and consistently measured, and current usage better understood; Project future <span class="hlt">energy</span> demand; Establish a system for centralized, ongoing tracking and analysis of tribal <span class="hlt">energy</span> data that is applicable across sectors, facilities, and activities; Develop a unifying vision that is consistent with the tribe’s long-term cultural, social, environmental, and economic goals; Identify and evaluate the potential of opportunities for development of long-term, cost effective <span class="hlt">energy</span> sources, such as renewable <span class="hlt">energy</span>, <span class="hlt">energy</span> efficiency and conservation, and other feasible supply- and demand-side options; and Build the SYBCI’s capacity for</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1226906-nudged-elastic-band-method-two-climbing-images-finding-transition-states-complex-energy-landscapes','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1226906-nudged-elastic-band-method-two-climbing-images-finding-transition-states-complex-energy-landscapes"><span>Nudged-elastic <span class="hlt">band</span> method with two climbing images: Finding transition states in complex <span class="hlt">energy</span> landscapes</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Zarkevich, Nikolai A.; Johnson, Duane D.</p> <p>2015-01-09</p> <p>The nudged-elastic <span class="hlt">band</span> (NEB) method is modified with concomitant two climbing images (C2-NEB) to find a transition state (TS) in complex <span class="hlt">energy</span> landscapes, such as those with a serpentine minimal <span class="hlt">energy</span> path (MEP). If a single climbing image (C1-NEB) successfully finds the TS, then C2-NEB finds it too. Improved stability of C2-NEB makes it suitable for more complex cases, where C1-NEB misses the TS because the MEP and NEB directions near the saddle point are different. Generally, C2-NEB not only finds the TS, but guarantees, by construction, that the climbing images approach it from the opposite sides along the MEP.more » In <span class="hlt">addition</span>, C2-NEB provides an accuracy estimate from the three images: the highest-<span class="hlt">energy</span> one and its climbing neighbors. C2-NEB is suitable for fixed-cell NEB and the generalized solid-state NEB.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1369394','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1369394"><span>Effects of an <span class="hlt">additional</span> conduction <span class="hlt">band</span> on the singlet-antiferromagnet competition in the periodic Anderson model</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Hu, Wenjian; Scalettar, Richard T.; Huang, Edwin W.</p> <p></p> <p>The competition between antiferromagnetic (AF) order and singlet formation is a central phenomenon of the Kondo and periodic Anderson Hamiltonians and of the heavy fermion materials they describe. In this paper, we explore the effects of an <span class="hlt">additional</span> conduction <span class="hlt">band</span> on magnetism in these models, and, specifically, on changes in the AF-singlet quantum critical point (QCP) and the one particle and spin spectral functions. To understand the magnetic phase transition qualitatively, we first carry out a self-consistent mean field theory (MFT). The basic conclusion is that, at half filling, the coupling to the <span class="hlt">additional</span> <span class="hlt">band</span> stabilizes the AF phase tomore » larger f d hybridization V in the PAM. We also explore the possibility of competing ferromagnetic phases when this conduction <span class="hlt">band</span> is doped away from half filling. Here, we next employ quantum Monte Carlo (QMC) which, in combination with finite size scaling, allows us to evaluate the position of the QCP using an exact treatment of the interactions. This approach confirms the stabilization of AF order, which occurs through an enhancement of the Ruderman-Kittel-Kasuya-Yosida (RKKY) interaction. QMC results for the spectral function A (q,ω) and dynamic spin structure factor χ (q,ω) yield <span class="hlt">additional</span> insight into the AF-singlet competition and the low temperature phases.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1369394-effects-additional-conduction-band-singlet-antiferromagnet-competition-periodic-anderson-model','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1369394-effects-additional-conduction-band-singlet-antiferromagnet-competition-periodic-anderson-model"><span>Effects of an <span class="hlt">additional</span> conduction <span class="hlt">band</span> on the singlet-antiferromagnet competition in the periodic Anderson model</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Hu, Wenjian; Scalettar, Richard T.; Huang, Edwin W.; ...</p> <p>2017-06-12</p> <p>The competition between antiferromagnetic (AF) order and singlet formation is a central phenomenon of the Kondo and periodic Anderson Hamiltonians and of the heavy fermion materials they describe. In this paper, we explore the effects of an <span class="hlt">additional</span> conduction <span class="hlt">band</span> on magnetism in these models, and, specifically, on changes in the AF-singlet quantum critical point (QCP) and the one particle and spin spectral functions. To understand the magnetic phase transition qualitatively, we first carry out a self-consistent mean field theory (MFT). The basic conclusion is that, at half filling, the coupling to the <span class="hlt">additional</span> <span class="hlt">band</span> stabilizes the AF phase tomore » larger f d hybridization V in the PAM. We also explore the possibility of competing ferromagnetic phases when this conduction <span class="hlt">band</span> is doped away from half filling. Here, we next employ quantum Monte Carlo (QMC) which, in combination with finite size scaling, allows us to evaluate the position of the QCP using an exact treatment of the interactions. This approach confirms the stabilization of AF order, which occurs through an enhancement of the Ruderman-Kittel-Kasuya-Yosida (RKKY) interaction. QMC results for the spectral function A (q,ω) and dynamic spin structure factor χ (q,ω) yield <span class="hlt">additional</span> insight into the AF-singlet competition and the low temperature phases.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26166580','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26166580"><span>Quantitative operando visualization of the <span class="hlt">energy</span> <span class="hlt">band</span> depth profile in solar cells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Chen, Qi; Mao, Lin; Li, Yaowen; Kong, Tao; Wu, Na; Ma, Changqi; Bai, Sai; Jin, Yizheng; Wu, Dan; Lu, Wei; Wang, Bing; Chen, Liwei</p> <p>2015-07-13</p> <p>The <span class="hlt">energy</span> <span class="hlt">band</span> alignment in solar cell devices is critically important because it largely governs elementary photovoltaic processes, such as the generation, separation, transport, recombination and collection of charge carriers. Despite the expenditure of considerable effort, the measurement of <span class="hlt">energy</span> <span class="hlt">band</span> depth profiles across multiple layers has been extremely challenging, especially for operando devices. Here we present direct visualization of the surface potential depth profile over the cross-sections of operando organic photovoltaic devices using scanning Kelvin probe microscopy. The convolution effect due to finite tip size and cantilever beam crosstalk has previously prohibited quantitative interpretation of scanning Kelvin probe microscopy-measured surface potential depth profiles. We develop a bias voltage-compensation method to address this critical problem and obtain quantitatively accurate measurements of the open-circuit voltage, built-in potential and electrode potential difference.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24783945','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24783945"><span>Direct imaging of <span class="hlt">band</span> profile in single layer MoS2 on graphite: quasiparticle <span class="hlt">energy</span> gap, metallic edge states, and edge <span class="hlt">band</span> bending.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zhang, Chendong; Johnson, Amber; Hsu, Chang-Lung; Li, Lain-Jong; Shih, Chih-Kang</p> <p>2014-05-14</p> <p>Using scanning tunneling microscopy and spectroscopy, we probe the electronic structures of single layer MoS2 on graphite. The apparent quasiparticle <span class="hlt">energy</span> gap of single layer MoS2 is measured to be 2.15 ± 0.06 eV at 77 K, albeit a higher second conduction <span class="hlt">band</span> threshold at 0.2 eV above the apparent conduction <span class="hlt">band</span> minimum is also observed. Combining it with photoluminescence studies, we deduce an exciton binding <span class="hlt">energy</span> of 0.22 ± 0.1 eV (or 0.42 eV if the second threshold is use), a value that is lower than current theoretical predictions. Consistent with theoretical predictions, we directly observe metallic edge states of single layer MoS2. In the bulk region of MoS2, the Fermi level is located at 1.8 eV above the valence <span class="hlt">band</span> maximum, possibly due to the formation of a graphite/MoS2 heterojunction. At the edge, however, we observe an upward <span class="hlt">band</span> bending of 0.6 eV within a short depletion length of about 5 nm, analogous to the phenomena of Fermi level pinning of a 3D semiconductor by metallic surface states.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2008JPCM...20g5233E','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2008JPCM...20g5233E"><span>Determination of the optical <span class="hlt">band</span>-gap <span class="hlt">energy</span> of cubic and hexagonal boron nitride using luminescence excitation spectroscopy</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Evans, D. A.; McGlynn, A. G.; Towlson, B. M.; Gunn, M.; Jones, D.; Jenkins, T. E.; Winter, R.; Poolton, N. R. J.</p> <p>2008-02-01</p> <p>Using synchrotron-based luminescence excitation spectroscopy in the <span class="hlt">energy</span> range 4-20 eV at 8 K, the indirect Γ-X optical <span class="hlt">band</span>-gap transition in cubic boron nitride is determined as 6.36 ± 0.03 eV, and the quasi-direct <span class="hlt">band</span>-gap <span class="hlt">energy</span> of hexagonal boron nitride is determined as 5.96 ± 0.04 eV. The composition and structure of the materials are self-consistently established by optically detected x-ray absorption spectroscopy, and both x-ray diffraction and Raman measurements on the same samples give independent confirmation of their chemical and structural purity: together, the results are therefore considered as providing definitive measurements of the optical <span class="hlt">band</span>-gap <span class="hlt">energies</span> of the two materials.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_1");'>1</a></li> <li><a href="#" onclick='return showDiv("page_2");'>2</a></li> <li class="active"><span>3</span></li> <li><a href="#" onclick='return showDiv("page_4");'>4</a></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_3 --> <div id="page_4" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_2");'>2</a></li> <li><a href="#" onclick='return showDiv("page_3");'>3</a></li> <li class="active"><span>4</span></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="61"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017E%26ES...75a2003W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017E%26ES...75a2003W"><span><span class="hlt">Band</span>-engineering of TiO2 as a wide-<span class="hlt">band</span> gap semiconductor using organic chromophore dyes</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wahyuningsih, S.; Kartini, I.; Ramelan, A. H.; Saputri, L. N. M. Z.; Munawaroh, H.</p> <p>2017-07-01</p> <p>Bond-engineering as applied to semiconductor materials refers to the manipulation of the <span class="hlt">energy</span> <span class="hlt">bands</span> in order to control charge transfer processes in a device. When the device in question is a photoelectrochemical cell, the charges affected by drift become the focus of the study. The ideal <span class="hlt">band</span> gap of semiconductors for enhancement of photocatalyst activity can be lowered to match with visible light absorption and the location of conduction <span class="hlt">Band</span> (CB) should be raised to meet the reducing capacity. Otherwise, by the <span class="hlt">addition</span> of the chromofor organic dyes, the wide-<span class="hlt">band</span> gab can be influences by interacation resulting between TiO2 surface and the dyes. We have done the impruvisation wide-<span class="hlt">band</span> gap of TiO2 by the <span class="hlt">addition</span> of organic chromophore dye, and the <span class="hlt">addition</span> of transition metal dopand. The TiO2 morphology influence the light absorption as well as the surface modification. The organic chromophore dye was syntesized by formation complexes compound of Co(PAR)(SiPA)(PAR)= 4-(2-piridylazoresorcinol), SiPA = Silyl propil amine). The result showed that the chromophore groups adsorbed onto TiO2 surface can increase the visible light absorption of wide-<span class="hlt">band</span> gab semiconductor. Initial absorption of a chromophore will affect light penetration into the material surfaces. The use of photonic material as a solar cell shows this phenomenon clearly from the IPCE (incident photon to current conversion efficiency) measurement data. Organic chromophore dyes of Co(PAR)(SiPA) exhibited the long wavelength absorption character compared to the N719 dye (from Dyesol).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017PhRvB..96t5206S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017PhRvB..96t5206S"><span>Simple vertex correction improves G W <span class="hlt">band</span> <span class="hlt">energies</span> of bulk and two-dimensional crystals</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Schmidt, Per S.; Patrick, Christopher E.; Thygesen, Kristian S.</p> <p>2017-11-01</p> <p>The G W self-<span class="hlt">energy</span> method has long been recognized as the gold standard for quasiparticle (QP) calculations of solids in spite of the fact that the neglect of vertex corrections and the use of a density-functional theory starting point lack rigorous justification. In this work we remedy this situation by including a simple vertex correction that is consistent with a local-density approximation starting point. We analyze the effect of the self-<span class="hlt">energy</span> by splitting it into short-range and long-range terms which are shown to govern, respectively, the center and size of the <span class="hlt">band</span> gap. The vertex mainly improves the short-range correlations and therefore has a small effect on the <span class="hlt">band</span> gap, while it shifts the <span class="hlt">band</span> gap center up in <span class="hlt">energy</span> by around 0.5 eV, in good agreement with experiments. Our analysis also explains how the relative importance of short- and long-range interactions in structures of different dimensionality is reflected in their QP <span class="hlt">energies</span>. Inclusion of the vertex comes at practically no extra computational cost and even improves the basis set convergence compared to G W . Taken together, the method provides an efficient and rigorous improvement over the G W approximation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/21513355-kev-band-linac-material-recognition-using-two-fold-scintillator-detector-concept-dual-energy-ray-system','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/21513355-kev-band-linac-material-recognition-using-two-fold-scintillator-detector-concept-dual-energy-ray-system"><span>950 keV X-<span class="hlt">Band</span> Linac For Material Recognition Using Two-Fold Scintillator Detector As A Concept Of Dual-<span class="hlt">Energy</span> X-Ray System</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Lee, Kiwoo; Natsui, Takuya; Hirai, Shunsuke</p> <p>2011-06-01</p> <p>One of the advantages of applying X-<span class="hlt">band</span> linear accelerator (Linac) is the compact size of the whole system. That shows us the possibility of on-site system such as the custom inspection system in an airport. As X-ray source, we have developed X-<span class="hlt">band</span> Linac and achieved maximum X-ray <span class="hlt">energy</span> 950 keV using the low power magnetron (250 kW) in 2 {mu}s pulse length. The whole size of the Linac system is 1x1x1 m{sup 3}. That is realized by introducing X-<span class="hlt">band</span> system. In <span class="hlt">addition</span>, we have designed two-fold scintillator detector in dual <span class="hlt">energy</span> X-ray concept. Monte carlo N-particle transport (MCNP) code wasmore » used to make up sensor part of the design with two scintillators, CsI and CdWO4. The custom inspection system is composed of two equipments: 950 keV X-<span class="hlt">band</span> Linac and two-fold scintillator and they are operated simulating real situation such as baggage check in an airport. We will show you the results of experiment which was performed with metal samples: iron and lead as targets in several conditions.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4510960','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4510960"><span>Quantitative operando visualization of the <span class="hlt">energy</span> <span class="hlt">band</span> depth profile in solar cells</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Chen, Qi; Mao, Lin; Li, Yaowen; Kong, Tao; Wu, Na; Ma, Changqi; Bai, Sai; Jin, Yizheng; Wu, Dan; Lu, Wei; Wang, Bing; Chen, Liwei</p> <p>2015-01-01</p> <p>The <span class="hlt">energy</span> <span class="hlt">band</span> alignment in solar cell devices is critically important because it largely governs elementary photovoltaic processes, such as the generation, separation, transport, recombination and collection of charge carriers. Despite the expenditure of considerable effort, the measurement of <span class="hlt">energy</span> <span class="hlt">band</span> depth profiles across multiple layers has been extremely challenging, especially for operando devices. Here we present direct visualization of the surface potential depth profile over the cross-sections of operando organic photovoltaic devices using scanning Kelvin probe microscopy. The convolution effect due to finite tip size and cantilever beam crosstalk has previously prohibited quantitative interpretation of scanning Kelvin probe microscopy-measured surface potential depth profiles. We develop a bias voltage-compensation method to address this critical problem and obtain quantitatively accurate measurements of the open-circuit voltage, built-in potential and electrode potential difference. PMID:26166580</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22025555-partially-filled-intermediate-band-cr-doped-gan-films','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22025555-partially-filled-intermediate-band-cr-doped-gan-films"><span>Partially filled intermediate <span class="hlt">band</span> of Cr-doped GaN films</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Sonoda, S.</p> <p>2012-05-14</p> <p>We investigated the <span class="hlt">band</span> structure of sputtered Cr-doped GaN (GaCrN) films using optical absorption, photoelectron yield spectroscopy, and charge transport measurements. It was found that an <span class="hlt">additional</span> <span class="hlt">energy</span> <span class="hlt">band</span> is formed in the intrinsic <span class="hlt">band</span> gap of GaN upon Cr doping, and that charge carriers in the material move in the inserted <span class="hlt">band</span>. Prototype solar cells showed enhanced short circuit current and open circuit voltage in the n-GaN/GaCrN/p-GaN structure compared to the GaCrN/p-GaN structure, which validates the proposed concept of an intermediate-<span class="hlt">band</span> solar cell.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016PhRvB..93h5202G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016PhRvB..93h5202G"><span>Quasiparticle <span class="hlt">band</span> gap of organic-inorganic hybrid perovskites: Crystal structure, spin-orbit coupling, and self-<span class="hlt">energy</span> effects</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gao, Weiwei; Gao, Xiang; Abtew, Tesfaye A.; Sun, Yi-Yang; Zhang, Shengbai; Zhang, Peihong</p> <p>2016-02-01</p> <p>The quasiparticle <span class="hlt">band</span> gap is one of the most important materials properties for photovoltaic applications. Often the <span class="hlt">band</span> gap of a photovoltaic material is determined (and can be controlled) by various factors, complicating predictive materials optimization. An in-depth understanding of how these factors affect the size of the gap will provide valuable guidance for new materials discovery. Here we report a comprehensive investigation on the <span class="hlt">band</span> gap formation mechanism in organic-inorganic hybrid perovskites by decoupling various contributing factors which ultimately determine their electronic structure and quasiparticle <span class="hlt">band</span> gap. Major factors, namely, quasiparticle self-<span class="hlt">energy</span>, spin-orbit coupling, and structural distortions due to the presence of organic molecules, and their influences on the quasiparticle <span class="hlt">band</span> structure of organic-inorganic hybrid perovskites are illustrated. We find that although methylammonium cations do not contribute directly to the electronic states near <span class="hlt">band</span> edges, they play an important role in defining the <span class="hlt">band</span> gap by introducing structural distortions and controlling the overall lattice constants. The spin-orbit coupling effects drastically reduce the electron and hole effective masses in these systems, which is beneficial for high carrier mobilities and small exciton binding <span class="hlt">energies</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JAP...121x4303Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JAP...121x4303Y"><span>Conductance modulation in Weyl semimetals with tilted <span class="hlt">energy</span> dispersion without a <span class="hlt">band</span> gap</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yesilyurt, Can; Siu, Zhuo Bin; Tan, Seng Ghee; Liang, Gengchiau; Jalil, Mansoor B. A.</p> <p>2017-06-01</p> <p>We investigate the tunneling conductance of Weyl semimetal with tilted <span class="hlt">energy</span> dispersion by considering electron transmission through a p-n-p junction with one-dimensional electric and magnetic barriers. In the presence of both electric and magnetic barriers, we found that a large conductance gap can be produced with the aid of tilted <span class="hlt">energy</span> dispersion without a <span class="hlt">band</span> gap. The origin of this effect is the shift of the electron wave-vector at barrier boundaries caused by (i) the pseudo-magnetic field induced by electrical potential, i.e., a newly discovered feature that is only possible in the materials possessing tilted <span class="hlt">energy</span> dispersion, (ii) the real magnetic field induced by a ferromagnetic layer deposited on the top of the system. We use a realistic barrier structure applicable in current nanotechnology and analyze the temperature dependence of the tunneling conductance. The new approach presented here may resolve a major problem of possible transistor applications in topological semimetals, i.e., the absence of normal backscattering and gapless <span class="hlt">band</span> structure.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016APS..MAR.G1313G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016APS..MAR.G1313G"><span>Measurement of the low <span class="hlt">energy</span> spectral contribution in coincidence with valence <span class="hlt">band</span> (VB) <span class="hlt">energy</span> levels of Ag(100) using VB-VB coincidence spectroscopy</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gladen, R. W.; Joglekar, P. V.; Lim, Z. H.; Shastry, K.; Hulbert, S. L.; Weiss, A. H.</p> <p></p> <p>A set of coincidence measurements were obtained for the study and measurement of the electron contribution arising from the inter-valence <span class="hlt">band</span> (VB) transitions along with the inelastically scattered VB electron contribution. These Auger-unrelated contributions arise in the Auger spectrum (Ag 4p NVV) obtained using Auger Photoelectron Coincidence Spectroscopy (APECS). The measured Auger-unrelated contribution can be eliminated from Auger spectrum to obtain the spectrum related to Auger. In our VB-VB coincidence measurement, a photon beam of <span class="hlt">energy</span> 180eV was used to probe the Ag(100) sample. The coincidence spectrum was obtained using two Cylindrical Mirror Analyzers (CMA's). The scan CMA measured the low <span class="hlt">energy</span> electron contribution in the <span class="hlt">energy</span> range 0-70eV in coincidence with VB electrons measured by the fixed CMA. In this talk, we present the data obtained for VB-VB coincidence at the valence <span class="hlt">band</span> <span class="hlt">energy</span> of 171eV along with the coincidence measurements in the <span class="hlt">energy</span> range of 4p core and valence <span class="hlt">band</span>. NSF DMR 0907679, NSF Award Number: 1213727. Use of the National Synchrotron Light Source, Brookhaven National Laboratory, was supported by the U.S. DOE, Office of Science, Office of Basic <span class="hlt">Energy</span> Sciences, under Contract No. DEAC02-98CH10886.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015APS..MARS21011J','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015APS..MARS21011J"><span>Measurement of the low <span class="hlt">energy</span> spectral contribution in coincidence with valence <span class="hlt">band</span> (VB) <span class="hlt">energy</span> levels of Ag(100) using VB-VB coincidence spectroscopy</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Joglekar, P. V.; Gladen, R.; Lim, Z. H.; Shastry, K.; Hulbert, S. L.; Weiss, A. H.</p> <p>2015-03-01</p> <p>A set of coincidence measurements were obtained for the study and measurement of the electron contribution arising from the inter-valence <span class="hlt">band</span> (VB) transitions along with the inelastically scattered VB electron contribution. These Auger-unrelated contributions arise in the Auger spectrum (Ag 4p NVV) obtained using Auger Photoelectron Coincidence Spectroscopy (APECS). The measured Auger-unrelated contribution can be eliminated from Auger spectrum to obtain the spectrum related to Auger. In our VB-VB coincidence measurement, a photon beam of <span class="hlt">energy</span> 180eV was used to probe the Ag(100) sample. The coincidence spectrum was obtained using two Cylindrical Mirror Analyzers (CMA's). The scan CMA measured the low <span class="hlt">energy</span> electron contribution in the <span class="hlt">energy</span> range 0-70eV in coincidence with VB electrons measured by the fixed CMA. In this talk, we present the data obtained for VB-VB coincidence at the valence <span class="hlt">band</span> <span class="hlt">energy</span> of 171eV along with the coincidence measurements in the <span class="hlt">energy</span> range of 4p core and valence <span class="hlt">band</span>. NSF DMR 0907679, NSF Award Number: 1213727. Use of the National Synchrotron Light Source, Brookhaven National Laboratory, was supported by the U.S. DOE, Office of Science, Office of Basic <span class="hlt">Energy</span> Sciences, under Contract No. DE-AC02-98CH10886.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012PhDT.......211M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012PhDT.......211M"><span><span class="hlt">Band</span> structure engineering for solar <span class="hlt">energy</span> applications: Zinc oxide(1-x) selenium(x) films and devices</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mayer, Marie Annette</p> <p></p> <p>New technologies motivate the development of new semiconducting materials, for which structural, electrical and chemical properties are not well understood. In <span class="hlt">addition</span> to new materials systems, there are huge opportunities for new applications, especially in solar <span class="hlt">energy</span> conversion. In this dissertation I explore the role of <span class="hlt">band</span> structure engineering of semiconducting oxides for solar <span class="hlt">energy</span>. Due to the abundance and electrochemical stability of oxides, the appropriate modification could make them appealing for applications in both photovoltaics and photoelectrochemical hydrogen production. This dissertation describes the design, synthesis and evaluation of the alloy ZnO1-xSe x for these purposes. I review several methods of <span class="hlt">band</span> structure engineering including strain, quantum confinement and alloying. A detailed description of the <span class="hlt">band</span> anticrossing (BAC) model for highly mismatched alloys is provided, including the derivation of the BAC model as well as recent work and potential applications. Thin film ZnOxSe1-x samples are grown by pulsed laser deposition (PLD). I describe in detail the effect of growth conditions (temperature, pressure and laser fluence) on the chemistry, structure and optoelectronic properties of ZnOxSe1-x. The films are grown using different combinations of PLD conditions and characterized with a variety of techniques. Phase pure films with low roughness and high crystallinity were obtained at temperatures below 450¢ªC, pressures less than 10-4 Torr and laser fluences on the order of 1.5 J/cm 2. Electrical conduction was still observed despite heavy concentrations of grain boundaries. The <span class="hlt">band</span> structure of ZnO1-xSex is then examined in detail. The bulk electron affinity of a ZnO thin film was measured to be 4.5 eV by pinning the Fermi level with native defects. This is explained in the framework of the amphoteric defect model. A shift in the ZnO1-xSe x valence <span class="hlt">band</span> edge with x is observed using synchrotron x-ray absorption and emission</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017CP....493..194Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017CP....493..194Z"><span>Plasmon enhanced heterogeneous electron transfer with continuous <span class="hlt">band</span> <span class="hlt">energy</span> model</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zhao, Dandan; Niu, Lu; Wang, Luxia</p> <p>2017-08-01</p> <p>Photoinduced charge injection from a perylene dye molecule into the conduction <span class="hlt">band</span> of a TiO2 system decorated by a metal nanoparticles (MNP) is studied theoretically. Utilizing the density matrix theory the charge transfer dynamics is analyzed. The continuous behavior of the TiO2 conduction <span class="hlt">band</span> is accounted for by a Legendre polynomials expansion. The simulations consider optical excitation of the dye molecule coupled to the MNP and the subsequent electron injection into the TiO2 semiconductor. Due to the <span class="hlt">energy</span> transfer coupling between the molecule and the MNP optical excitation and subsequent charge injection into semiconductor is strongly enhanced. The respective enhancement factor can reach values larger than 103. Effects of pulse duration, coupling strength and energetic resonances are also analyzed. The whole approach offers an efficient way to increase charge injection in dye-sensitized solar cells.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017SeScT..32j4008F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017SeScT..32j4008F"><span>Control of Ge1-x-ySixSny layer lattice constant for <span class="hlt">energy</span> <span class="hlt">band</span> alignment in Ge1-xSnx/Ge1-x-ySixSny heterostructures</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Fukuda, Masahiro; Watanabe, Kazuhiro; Sakashita, Mitsuo; Kurosawa, Masashi; Nakatsuka, Osamu; Zaima, Shigeaki</p> <p>2017-10-01</p> <p>The <span class="hlt">energy</span> <span class="hlt">band</span> alignment of Ge1-xSnx/Ge1-x-ySixSny heterostructures was investigated, and control of the valence <span class="hlt">band</span> offset at the Ge1-xSnx/Ge1-x-ySixSny heterointerface was achieved by controlling the Si and Sn contents in the Ge1-x-ySixSny layer. The valence <span class="hlt">band</span> offset in the Ge0.902Sn0.098/Ge0.41Si0.50Sn0.09 heterostructure was evaluated to be as high as 330 meV, and its conduction <span class="hlt">band</span> offset was estimated to be 150 meV by considering the <span class="hlt">energy</span> bandgap calculated from the theoretical prediction. In <span class="hlt">addition</span>, the formation of the strain-relaxed Ge1-x-ySixSny layer was examined and the crystalline structure was characterized. The epitaxial growth of a strain-relaxed Ge0.64Si0.21Sn0.15 layer with the degree of strain relaxation of 55% was examined using a virtual Ge substrate. Moreover, enhancement of the strain relaxation was demonstrated by post-deposition annealing, where a degree of strain relaxation of 70% was achieved after annealing at 400 °C. These results indicate the possibility for enhancing the indirect-direct crossover with a strained and high-Sn-content Ge1-xSnx layer on a strain-relaxed Ge1-x-ySixSny layer, realizing preferable carrier confinement by type-I <span class="hlt">energy</span> <span class="hlt">band</span> alignment with high conduction and valence <span class="hlt">band</span> offsets.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28421210','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28421210"><span>More accurate depiction of adsorption <span class="hlt">energy</span> on transition metals using work function as one <span class="hlt">additional</span> descriptor.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Shen, Xiaochen; Pan, Yanbo; Liu, Bin; Yang, Jinlong; Zeng, Jie; Peng, Zhenmeng</p> <p>2017-05-24</p> <p>The reaction mechanism and properties of a catalytic process are primarily determined by the interactions between reacting species and catalysts. However, the interactions are often challenging to be experimentally measured, especially for unstable intermediates. Therefore, it is of significant importance to establish an exact relationship between chemical-catalyst interactions and catalyst parameters, which will allow calculation of these interactions and thus advance their mechanistic understanding. Herein we report the description of adsorption <span class="hlt">energy</span> on transition metals by considering both ionic bonding and covalent bonding contributions and introduce the work function as one <span class="hlt">additional</span> responsible parameter. We find that the adsorption <span class="hlt">energy</span> can be more accurately described using a two-dimensional (2D) polynomial model, which shows a significant improvement compared with the current adsorption <span class="hlt">energy-d-band</span> center linear correlation. We also demonstrate the utilization of this new 2D polynomial model to calculate oxygen binding <span class="hlt">energy</span> of different transition metals to help understand their catalytic properties in oxygen reduction reactions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JPhD...51u5102B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JPhD...51u5102B"><span><span class="hlt">Band</span>-edges and <span class="hlt">band</span>-gap in few-layered transition metal dichalcogenides</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bhunia, Hrishikesh; Pal, Amlan J.</p> <p>2018-05-01</p> <p>We have considered liquid-exfoliated transition metal dichalcogenides (WS2, WSe2, MoS2, and MoSe2) and studied their <span class="hlt">band</span>-edges and <span class="hlt">band</span>-gap through scanning tunneling spectroscopy (STS) and density of states. A monolayer, bilayer (2L), and trilayer (3L) of each of the layered materials were characterized to derive the <span class="hlt">energies</span>. Upon an increase in the number of layers, both the <span class="hlt">band</span>-edges were found to shift towards the Fermi <span class="hlt">energy</span>. The results from the exfoliated nanosheets have been compared with reported STS studies of MoS2 and WSe2 formed through chemical vapor deposition or molecular beam epitaxy methods; an uncontrolled lattice strain existed in such 2L and 3L nanoflakes due to mismatch in stacking-patterns between the monolayers affecting their <span class="hlt">energies</span>. In the present work, the layers formed through the liquid-exfoliation process retained their interlayer coupling or stacking-sequence prevalent to the bulk and hence allowed determination of <span class="hlt">band-energies</span> in these strain-free two-dimensional materials.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JLTP..191...49P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JLTP..191...49P"><span>Quantum Transport and Non-Hermiticity on Flat-<span class="hlt">Band</span> Lattices</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Park, Hee Chul; Ryu, Jung-Wan; Myoung, Nojoon</p> <p>2018-04-01</p> <p>We investigate quantum transport in a flat-<span class="hlt">band</span> lattice induced in a twisted cross-stitch lattice with Hermitian or non-Hermitian potentials, with a combination of parity and time-reversal symmetry invariant. In the given system, the transmission probability demonstrates a resonant behavior on the real part of the <span class="hlt">energy</span> <span class="hlt">bands</span>. Both of the potentials break the parity symmetry, which lifts the degeneracy of the flat and dispersive <span class="hlt">bands</span>. In <span class="hlt">addition</span>, non-Hermiticity conserving PT-symmetry induces a transition between the unbroken and broken PT-symmetric phases through exceptional points in momentum space. Characteristics of non-Hermitian and Hermitian bandgaps are distinguishable: The non-Hermitian bandgap is induced by separation toward complex <span class="hlt">energy</span>, while the Hermitian bandgap is caused by the expelling of available states into real <span class="hlt">energy</span>. Deviation of the two bandgaps follows as a function of the quartic power of the induced potential. It is notable that non-Hermiticity plays an important role in the mechanism of generating a bandgap distinguishable from a Hermitian bandgap.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27812035','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27812035"><span>Engineering of <span class="hlt">band</span> gap states of amorphous SiZnSnO semiconductor as a function of Si doping concentration.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Choi, Jun Young; Heo, Keun; Cho, Kyung-Sang; Hwang, Sung Woo; Kim, Sangsig; Lee, Sang Yeol</p> <p>2016-11-04</p> <p>We investigated the <span class="hlt">band</span> gap of SiZnSnO (SZTO) with different Si contents. <span class="hlt">Band</span> gap engineering of SZTO is explained by the evolution of the electronic structure, such as changes in the <span class="hlt">band</span> edge states and <span class="hlt">band</span> gap. Using ultraviolet photoelectron spectroscopy (UPS), it was verified that Si atoms can modify the <span class="hlt">band</span> gap of SZTO thin films. Carrier generation originating from oxygen vacancies can modify the <span class="hlt">band</span>-gap states of oxide films with the <span class="hlt">addition</span> of Si. Since it is not easy to directly derive changes in the <span class="hlt">band</span> gap states of amorphous oxide semiconductors, no reports of the relationship between the Fermi <span class="hlt">energy</span> level of oxide semiconductor and the device stability of oxide thin film transistors (TFTs) have been presented. The <span class="hlt">addition</span> of Si can reduce the total density of trap states and change the <span class="hlt">band</span>-gap properties. When 0.5 wt% Si was used to fabricate SZTO TFTs, they showed superior stability under negative bias temperature stress. We derived the <span class="hlt">band</span> gap and Fermi <span class="hlt">energy</span> level directly using data from UPS, Kelvin probe, and high-resolution electron <span class="hlt">energy</span> loss spectroscopy analyses.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5095643','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5095643"><span>Engineering of <span class="hlt">band</span> gap states of amorphous SiZnSnO semiconductor as a function of Si doping concentration</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Choi, Jun Young; Heo, Keun; Cho, Kyung-Sang; Hwang, Sung Woo; Kim, Sangsig; Lee, Sang Yeol</p> <p>2016-01-01</p> <p>We investigated the <span class="hlt">band</span> gap of SiZnSnO (SZTO) with different Si contents. <span class="hlt">Band</span> gap engineering of SZTO is explained by the evolution of the electronic structure, such as changes in the <span class="hlt">band</span> edge states and <span class="hlt">band</span> gap. Using ultraviolet photoelectron spectroscopy (UPS), it was verified that Si atoms can modify the <span class="hlt">band</span> gap of SZTO thin films. Carrier generation originating from oxygen vacancies can modify the <span class="hlt">band</span>-gap states of oxide films with the <span class="hlt">addition</span> of Si. Since it is not easy to directly derive changes in the <span class="hlt">band</span> gap states of amorphous oxide semiconductors, no reports of the relationship between the Fermi <span class="hlt">energy</span> level of oxide semiconductor and the device stability of oxide thin film transistors (TFTs) have been presented. The <span class="hlt">addition</span> of Si can reduce the total density of trap states and change the <span class="hlt">band</span>-gap properties. When 0.5 wt% Si was used to fabricate SZTO TFTs, they showed superior stability under negative bias temperature stress. We derived the <span class="hlt">band</span> gap and Fermi <span class="hlt">energy</span> level directly using data from UPS, Kelvin probe, and high-resolution electron <span class="hlt">energy</span> loss spectroscopy analyses. PMID:27812035</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JaJAP..57fKA05Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JaJAP..57fKA05Y"><span><span class="hlt">Energy</span> <span class="hlt">band</span> structure and electrical properties of Ga-oxide/GaN interface formed by remote oxygen plasma</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yamamoto, Taishi; Taoka, Noriyuki; Ohta, Akio; Truyen, Nguyen Xuan; Yamada, Hisashi; Takahashi, Tokio; Ikeda, Mitsuhisa; Makihara, Katsunori; Nakatsuka, Osamu; Shimizu, Mitsuaki; Miyazaki, Seiichi</p> <p>2018-06-01</p> <p>The <span class="hlt">energy</span> <span class="hlt">band</span> structure of a Ga-oxide/GaN structure formed by remote oxygen plasma exposure and the electrical interface properties of the GaN metal–oxide–semiconductor (MOS) capacitors with the SiO2/Ga-oxide/GaN structures with postdeposition annealing (PDA) at various temperatures have been investigated. Reflection high-<span class="hlt">energy</span> electron diffraction and X-ray photoelectron spectroscopy clarified that the formed Ga-oxide layer is neither a single nor polycrystalline phase with high crystallinity. We found that the <span class="hlt">energy</span> <span class="hlt">band</span> offsets at the conduction <span class="hlt">band</span> minimum and at the valence <span class="hlt">band</span> maximum between the Ga-oxide layer and the GaN surface were 0.4 and 1.2 ± 0.2 eV, respectively. Furthermore, capacitance–voltage (C–V) characteristics revealed that the interface trap density (D it) is lower than the evaluation limit of Terman method without depending on the PDA temperatures, and that the SiO2/Ga-oxide stack can work as a protection layer to maintain the low D it, avoiding the significant decomposition of GaN at the high PDA temperature of 800 °C.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012AIPA....2b2111C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012AIPA....2b2111C"><span>Lateral <span class="hlt">energy</span> <span class="hlt">band</span> profile modulation in tunnel field effect transistors based on gate structure engineering</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Cui, Ning; Liang, Renrong; Wang, Jing; Xu, Jun</p> <p>2012-06-01</p> <p>Choosing novel materials and structures is important for enhancing the on-state current in tunnel field-effect transistors (TFETs). In this paper, we reveal that the on-state performance of TFETs is mainly determined by the <span class="hlt">energy</span> <span class="hlt">band</span> profile of the channel. According to this interpretation, we present a new concept of <span class="hlt">energy</span> <span class="hlt">band</span> profile modulation (BPM) achieved with gate structure engineering. It is believed that this approach can be used to suppress the ambipolar effect. Based on this method, a Si TFET device with a symmetrical tri-material-gate (TMG) structure is proposed. Two-dimensional numerical simulations demonstrated that the special <span class="hlt">band</span> profile in this device can boost on-state performance, and it also suppresses the off-state current induced by the ambipolar effect. These unique advantages are maintained over a wide range of gate lengths and supply voltages. The BPM concept can serve as a guideline for improving the performance of nanoscale TFET devices.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2000STIN...0171553M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2000STIN...0171553M"><span>Mars Global Surveyor Ka-<span class="hlt">Band</span> Frequency Data Analysis</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Morabito, D.; Butman, S.; Shambayati, S.</p> <p>2000-01-01</p> <p> for the feed and electronics equipment. A dichroic plate is used to reflect the X-<span class="hlt">band</span> <span class="hlt">energy</span> and pass the Ka-<span class="hlt">band</span> <span class="hlt">energy</span> to another mirror. The RF <span class="hlt">energy</span> for each <span class="hlt">band</span> is then focused onto a feed horn and low-noise amplifier package. After amplification and RF/IF downconversion, the IF signals are sent to the Experimental Tone Tracker (ETT), a digital phase-lock-loop receiver, which simultaneously tracks both X-<span class="hlt">band</span> and Ka-<span class="hlt">band</span> carrier signals. Once a signal is detected, the ETT outputs estimates of the SNR in a I -Hz bandwidth (Pc/No), baseband phase and frequency of the signals every I -sec. Between December 1996 and December 1998, the Ka-<span class="hlt">band</span> and X-<span class="hlt">band</span> signals from MGS were tracked on a regular basis using the ETT. The Ka-<span class="hlt">band</span> downlink frequencies described here were referenced to the spacecraft's on-board USO which was also the X-<span class="hlt">band</span> frequency reference (fka= 3.8 fx). The ETT estimates of baseband phase at I -second sampled time tags were converted to sky frequency estimates. Frequency residuals were then generated for each <span class="hlt">band</span> by removing a model frequency from each observable frequency at each time tag. The model included Doppler and other effects derived from spacecraft trajectory files obtained from the MGS Navigation Team. A simple troposphere correction was applied to the data. In <span class="hlt">addition</span> to residuals, the USO frequencies emitted by the spacecraft were estimated. For several passes, the USO frequencies were determined from X-<span class="hlt">band</span> data and from Ka-<span class="hlt">band</span> data (referred to X-<span class="hlt">band</span> by dividing by 3.8) and were found to be in good agreement. In <span class="hlt">addition</span>, X-<span class="hlt">band</span> USO frequency estimates from MGS Radio Science data acquired from operational DSN stations were available for comparison and were found to agree within the I Hz level. The remaining sub-Hertz differences were attributed to the different models and software algorithms used by MGS Radio Science and KaBLE-11. A summary of the results of a linear fit of the USO frequency versus time (day of year) is</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_2");'>2</a></li> <li><a href="#" onclick='return showDiv("page_3");'>3</a></li> <li class="active"><span>4</span></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_4 --> <div id="page_5" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_3");'>3</a></li> <li><a href="#" onclick='return showDiv("page_4");'>4</a></li> <li class="active"><span>5</span></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="81"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29460856','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29460856"><span>Ultraviolet photoelectron spectroscopy reveals <span class="hlt">energy-band</span> dispersion for π-stacked 7,8,15,16-tetraazaterrylene thin films in a donor-acceptor bulk heterojunction.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Aghdassi, Nabi; Wang, Qi; Ji, Ru-Ru; Wang, Bin; Fan, Jian; Duhm, Steffen</p> <p>2018-05-11</p> <p>7,8,15,16-tetraazaterrylene (TAT) thin films grown on highly oriented pyrolytic graphite (HOPG) substrates were studied extensively with regard to their intrinsic and interfacial electronic properties by means of ultraviolet photoelectron spectroscopy (UPS). Merely weak substrate-adsorbate interaction occurs at the TAT/HOPG interface, with interface energetics being only little affected by the nominal film thickness. Photon <span class="hlt">energy</span>-dependent UPS performed perpendicular to the molecular planes of TAT multilayer films at room temperature clearly reveals <span class="hlt">band</span>-like intermolecular dispersion of the TAT highest occupied molecular orbital (HOMO) <span class="hlt">energy</span>. Based on a comparison with a tight-binding model, a relatively narrow bandwidth of 54 meV is derived, which points to the presence of an intermediate regime between hopping and <span class="hlt">band</span>-like hole transport. Upon <span class="hlt">additional</span> deposition of 2,2':5',2″:5″,2″'-quaterthiophene (4T), a 4T:TAT donor-acceptor bulk heterojunction with a considerable HOMO-level offset at the donor-acceptor interface is formed. The 4T:TAT bulk heterojunction likewise exhibits intermolecular dispersion of the TAT HOMO <span class="hlt">energy</span>, yet with a significant decreased bandwidth.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22596765-combined-analysis-energy-band-diagram-equivalent-circuit-nanocrystal-solid','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22596765-combined-analysis-energy-band-diagram-equivalent-circuit-nanocrystal-solid"><span>Combined analysis of <span class="hlt">energy</span> <span class="hlt">band</span> diagram and equivalent circuit on nanocrystal solid</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Kano, Shinya, E-mail: kano@eedept.kobe-u.ac.jp, E-mail: fujii@eedept.kobe-u.ac.jp; Sasaki, Masato; Fujii, Minoru, E-mail: kano@eedept.kobe-u.ac.jp, E-mail: fujii@eedept.kobe-u.ac.jp</p> <p></p> <p>We investigate a combined analysis of an <span class="hlt">energy</span> <span class="hlt">band</span> diagram and an equivalent circuit on nanocrystal (NC) solids. We prepared a flat silicon-NC solid in order to carry out the analysis. An <span class="hlt">energy</span> <span class="hlt">band</span> diagram of a NC solid is determined from DC transport properties. Current-voltage characteristics, photocurrent measurements, and conductive atomic force microscopy images indicate that a tunneling transport through a NC solid is dominant. Impedance spectroscopy gives an equivalent circuit: a series of parallel resistor-capacitors corresponding to NC/metal and NC/NC interfaces. The equivalent circuit also provides an evidence that the NC/NC interface mainly dominates the carrier transport throughmore » NC solids. Tunneling barriers inside a NC solid can be taken into account in a combined capacitance. Evaluated circuit parameters coincide with simple geometrical models of capacitances. As a result, impedance spectroscopy is also a useful technique to analyze semiconductor NC solids as well as usual DC transport. The analyses provide indispensable information to implement NC solids into actual electronic devices.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018MNRAS.476.2717H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018MNRAS.476.2717H"><span>A novel approach for characterizing broad-<span class="hlt">band</span> radio spectral <span class="hlt">energy</span> distributions</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Harvey, V. M.; Franzen, T.; Morgan, J.; Seymour, N.</p> <p>2018-05-01</p> <p>We present a new broad-<span class="hlt">band</span> radio frequency catalogue across 0.12 GHz ≤ ν ≤ 20 GHz created by combining data from the Murchison Widefield Array Commissioning Survey, the Australia Telescope 20 GHz survey, and the literature. Our catalogue consists of 1285 sources limited by S20 GHz > 40 mJy at 5σ, and contains flux density measurements (or estimates) and uncertainties at 0.074, 0.080, 0.119, 0.150, 0.180, 0.408, 0.843, 1.4, 4.8, 8.6, and 20 GHz. We fit a second-order polynomial in log-log space to the spectral <span class="hlt">energy</span> distributions of all these sources in order to characterize their broad-<span class="hlt">band</span> emission. For the 994 sources that are well described by a linear or quadratic model we present a new diagnostic plot arranging sources by the linear and curvature terms. We demonstrate the advantages of such a plot over the traditional radio colour-colour diagram. We also present astrophysical descriptions of the sources found in each segment of this new parameter space and discuss the utility of these plots in the upcoming era of large area, deep, broad-<span class="hlt">band</span> radio surveys.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/1388396-control-valence-conduction-band-energies-layered-transition-metal-phosphates-via-surface-functionalization','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/1388396-control-valence-conduction-band-energies-layered-transition-metal-phosphates-via-surface-functionalization"><span>Control of valence and conduction <span class="hlt">band</span> <span class="hlt">energies</span> in layered transition metal phosphates via surface functionalization</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Lentz, Levi C.; Kolb, Brian; Kolpak, Alexie M.</p> <p></p> <p>Layered transition metal phosphates and phosphites (TMPs) are a class of 2D materials bound togetherviavan der Waals interactions. Through simple functionalization, <span class="hlt">band</span> <span class="hlt">energies</span> can be systematically controlled.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28657448','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28657448"><span>Nudged elastic <span class="hlt">band</span> method and density functional theory calculation for finding a local minimum <span class="hlt">energy</span> pathway of p-benzoquinone and phenol fragmentation in mass spectrometry.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Sugimura, Natsuhiko; Igarashi, Yoko; Aoyama, Reiko; Shibue, Toshimichi</p> <p>2017-02-01</p> <p>Analysis of the fragmentation pathways of molecules in mass spectrometry gives a fundamental insight into gas-phase ion chemistry. However, the conventional intrinsic reaction coordinates method requires knowledge of the transition states of ion structures in the fragmentation pathways. Herein, we use the nudged elastic <span class="hlt">band</span> method, using only the initial and final state ion structures in the fragmentation pathways, and report the advantages and limitations of the method. We found a minimum <span class="hlt">energy</span> path of p-benzoquinone ion fragmentation with two saddle points and one intermediate structure. The primary <span class="hlt">energy</span> barrier, which corresponded to the cleavage of the C-C bond adjacent to the CO group, was calculated to be 1.50 eV. An <span class="hlt">additional</span> <span class="hlt">energy</span> barrier, which corresponded to the cleavage of the CO group, was calculated to be 0.68 eV. We also found an <span class="hlt">energy</span> barrier of 3.00 eV, which was the rate determining step of the keto-enol tautomerization in CO elimination from the molecular ion of phenol. The nudged elastic <span class="hlt">band</span> method allowed the determination of a minimum <span class="hlt">energy</span> path using only the initial and final state ion structures in the fragmentation pathways, and it provided faster than the conventional intrinsic reaction coordinates method. In <span class="hlt">addition</span>, this method was found to be effective in the analysis of the charge structures of the molecules during the fragmentation in mass spectrometry.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/934737','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/934737"><span>Augustine <span class="hlt">Band</span> of Cahuilla Indians <span class="hlt">Energy</span> Conservation and Options Analysis - Final Report</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Paul Turner</p> <p>2008-07-11</p> <p>The Augustine <span class="hlt">Band</span> of Cahuilla Indians was awarded a grant through the Department of <span class="hlt">Energy</span> First Steps program in June of 2006. The primary purpose of the grant was to enable the Tribe to develop <span class="hlt">energy</span> conservation policies and a strategy for alternative <span class="hlt">energy</span> resource development. All of the work contemplated by the grant agreement has been completed and the Tribe has begun implementing the resource development strategy through the construction of a 1.0 MW grid-connected photovoltaic system designed to offset a portion of the <span class="hlt">energy</span> demand generated by current and projected land uses on the Tribe’s Reservation. Implementation ofmore » proposed <span class="hlt">energy</span> conservation policies will proceed more deliberately as the Tribe acquires economic development experience sufficient to evaluate more systematically the interrelationships between conservation and its economic development goals.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22493112-detailed-analysis-energy-levels-configuration-existing-band-gap-supersaturated-silicon-titanium-photovoltaic-applications','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22493112-detailed-analysis-energy-levels-configuration-existing-band-gap-supersaturated-silicon-titanium-photovoltaic-applications"><span>A detailed analysis of the <span class="hlt">energy</span> levels configuration existing in the <span class="hlt">band</span> gap of supersaturated silicon with titanium for photovoltaic applications</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Pérez, E.; Dueñas, S.; Castán, H.</p> <p>2015-12-28</p> <p>The <span class="hlt">energy</span> levels created in supersaturated n-type silicon substrates with titanium implantation in the attempt to create an intermediate <span class="hlt">band</span> in their <span class="hlt">band</span>-gap are studied in detail. Two titanium ion implantation doses (10{sup 13 }cm{sup -2} and 10{sup 14 }cm{sup -2}) are studied in this work by conductance transient technique and admittance spectroscopy. Conductance transients have been measured at temperatures of around 100 K. The particular shape of these transients is due to the formation of <span class="hlt">energy</span> barriers in the conduction <span class="hlt">band</span>, as a consequence of the <span class="hlt">band</span>-gap narrowing induced by the high titanium concentration. Moreover, stationary admittance spectroscopy results suggest the existencemore » of different <span class="hlt">energy</span> level configuration, depending on the local titanium concentration. A continuum <span class="hlt">energy</span> level <span class="hlt">band</span> is formed when titanium concentration is over the Mott limit. On the other hand, when titanium concentration is lower than the Mott limit, but much higher than the donor impurity density, a quasi-continuum <span class="hlt">energy</span> level distribution appears. Finally, a single deep center appears for low titanium concentration. At the n-type substrate, the experimental results obtained by means of thermal admittance spectroscopy at high reverse bias reveal the presence of single levels located at around E{sub c}-425 and E{sub c}-275 meV for implantation doses of 10{sup 13 }cm{sup −2} and 10{sup 14 }cm{sup −2}, respectively. At low reverse bias voltage, quasi-continuously distributed <span class="hlt">energy</span> levels between the minimum of the conduction <span class="hlt">bands</span>, E{sub c} and E{sub c}-450 meV, are obtained for both doses. Conductance transients detected at low temperatures reveal that the high impurity concentration induces a <span class="hlt">band</span> gap narrowing which leads to the formation of a barrier in the conduction <span class="hlt">band</span>. Besides, the relationship between the activation <span class="hlt">energy</span> and the capture cross section values of all the <span class="hlt">energy</span> levels fits very well to the Meyer-Neldel rule. As it is</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015MS%26E...73a2100V','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015MS%26E...73a2100V"><span>Effect of Γ-X <span class="hlt">band</span> mixing on the donor binding <span class="hlt">energy</span> in a Quantum Wire</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Vijaya Shanthi, R.; Jayakumar, K.; Nithiananthi, P.</p> <p>2015-02-01</p> <p>To invoke the technological applications of heterostructure semiconductors like Quantum Well (QW), Quantum Well Wire (QWW) and Quantum Dot (QD), it is important to understand the property of impurity <span class="hlt">energy</span> which is responsible for the peculiar electronic & optical behavior of the Low Dimensional Semiconductor Systems (LDSS). Application of hydrostatic pressure P>35kbar drastically alters the <span class="hlt">band</span> offsets leading to the crossover of Γ <span class="hlt">band</span> of the well & X <span class="hlt">band</span> of the barrier resulting in an indirect transition of the carrier and this effect has been studied experimentally and theoretically in a QW structure. In this paper, we have investigated the effect of Γ-X <span class="hlt">band</span> mixing due to the application of hydrostatic pressure in a GaAs/AlxGa1-xAs QWW system. The results are presented and discussed for various widths of the wire.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014ApPhL.105k2108Q','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014ApPhL.105k2108Q"><span><span class="hlt">Band</span> gap and electronic structure of MgSiN2</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Quirk, J. B.; Râsander, M.; McGilvery, C. M.; Palgrave, R.; Moram, M. A.</p> <p>2014-09-01</p> <p>Density functional theory calculations and electron <span class="hlt">energy</span> loss spectroscopy indicate that the electronic structure of ordered orthorhombic MgSiN2 is similar to that of wurtzite AlN. A <span class="hlt">band</span> gap of 5.7 eV was calculated for both MgSiN2 (indirect) and AlN (direct) using the Heyd-Scuseria-Ernzerhof approximation. Correction with respect to the experimental room-temperature <span class="hlt">band</span> gap of AlN indicates that the true <span class="hlt">band</span> gap of MgSiN2 is 6.2 eV. MgSiN2 has an <span class="hlt">additional</span> direct gap of 6.3 eV at the Γ point.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018MPLA...3350048S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018MPLA...3350048S"><span><span class="hlt">Band</span> head spin assignment of superdeformed <span class="hlt">bands</span> in 133Pr using two-parameter formulae</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sharma, Honey; Mittal, H. M.</p> <p>2018-03-01</p> <p>The two-parameter formulae viz. the power index formula, the nuclear softness formula and the VMI model are adopted to accredit the <span class="hlt">band</span> head spin (I0) of four superdeformed rotational <span class="hlt">bands</span> in 133Pr. The technique of least square fitting is used to accredit the <span class="hlt">band</span> head spin for four superdeformed rotational <span class="hlt">bands</span> in 133Pr. The root mean deviation among the computed transition <span class="hlt">energies</span> and well-known experimental transition <span class="hlt">energies</span> are attained by extracting the model parameters from the two-parameter formulae. The determined transition <span class="hlt">energies</span> are in excellent agreement with the experimental transition <span class="hlt">energies</span>, whenever exact spins are accredited. The power index formula coincides well with the experimental data and provides minimum root mean deviation. So, the power index formula is more efficient tool than the nuclear softness formula and the VMI model. The deviation of dynamic moment of inertia J(2) against the rotational frequency is also examined.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20150011873','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20150011873"><span>A Novel Ku-<span class="hlt">Band/Ka-Band</span> and Ka-<span class="hlt">Band/E-Band</span> Multimode Waveguide Couplers for Power Measurement of Traveling-Wave Tube Amplifier Harmonic Frequencies</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Wintucky, Edwin G.; Simons, Rainee N.</p> <p>2015-01-01</p> <p>This paper presents the design, fabrication and test results for a novel waveguide multimode directional coupler (MDC). The coupler, fabricated from two dissimilar frequency <span class="hlt">band</span> waveguides, is capable of isolating power at the second harmonic frequency from the fundamental power at the output port of a traveling-wave tube (TWT) amplifier. Test results from proof-of-concept demonstrations are presented for a Ku-<span class="hlt">band/Ka-band</span> MDC and a Ka-<span class="hlt">band/E-band</span> MDC. In <span class="hlt">addition</span> to power measurements at harmonic frequencies, a potential application of the MDC is in the design of a satellite borne beacon source for atmospheric propagation studies at millimeter-wave (mm-wave) frequencies (Ka-<span class="hlt">band</span> and E-<span class="hlt">band</span>).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014APS..MARB44008S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014APS..MARB44008S"><span>Relating the defect <span class="hlt">band</span> gap and the density functional <span class="hlt">band</span> gap</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Schultz, Peter; Edwards, Arthur</p> <p>2014-03-01</p> <p>Density functional theory (DFT) is an important tool to probe the physics of materials. The Kohn-Sham (KS) gap in DFT is typically (much) smaller than the observed <span class="hlt">band</span> gap for materials in nature, the infamous ``<span class="hlt">band</span> gap problem.'' Accurate prediction of defect <span class="hlt">energy</span> levels is often claimed to be a casualty--the <span class="hlt">band</span> gap defines the <span class="hlt">energy</span> scale for defect levels. By applying rigorous control of boundary conditions in size-converged supercell calculations, however, we compute defect levels in Si and GaAs with accuracies of ~0.1 eV, across the full gap, unhampered by a <span class="hlt">band</span> gap problem. Using GaAs as a theoretical laboratory, we show that the defect <span class="hlt">band</span> gap--the span of computed defect levels--is insensitive to variations in the KS gap (with functional and pseudopotential), these KS gaps ranging from 0.1 to 1.1 eV. The defect gap matches the experimental 1.52 eV gap. The computed defect gaps for several other III-V, II-VI, I-VII, and other compounds also agree with the experimental gap, and show no correlation with the KS gap. Where, then, is the <span class="hlt">band</span> gap problem? This talk presents these results, discusses why the defect gap and the KS gap are distinct, implying that current understanding of what the ``<span class="hlt">band</span> gap problem'' means--and how to ``fix'' it--need to be rethought. Sandia National Laboratories is a multi-program laboratory managed and operated by Sandia Corporation, a wholly owned subsidiary of Lockheed Martin Company, for the U.S. Department of <span class="hlt">Energy</span>'s NNSA under contract DE-AC04-94AL85000.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018PhRvL.120n6402S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PhRvL.120n6402S"><span>Topological <span class="hlt">Band</span> Theory for Non-Hermitian Hamiltonians</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Shen, Huitao; Zhen, Bo; Fu, Liang</p> <p>2018-04-01</p> <p>We develop the topological <span class="hlt">band</span> theory for systems described by non-Hermitian Hamiltonians, whose <span class="hlt">energy</span> spectra are generally complex. After generalizing the notion of gapped <span class="hlt">band</span> structures to the non-Hermitian case, we classify "gapped" <span class="hlt">bands</span> in one and two dimensions by explicitly finding their topological invariants. We find nontrivial generalizations of the Chern number in two dimensions, and a new classification in one dimension, whose topology is determined by the <span class="hlt">energy</span> dispersion rather than the <span class="hlt">energy</span> eigenstates. We then study the bulk-edge correspondence and the topological phase transition in two dimensions. Different from the Hermitian case, the transition generically involves an extended intermediate phase with complex-<span class="hlt">energy</span> <span class="hlt">band</span> degeneracies at isolated "exceptional points" in momentum space. We also systematically classify all types of <span class="hlt">band</span> degeneracies.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017EPJWC.14612031M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EPJWC.14612031M"><span>Optical model with multiple <span class="hlt">band</span> couplings using soft rotator structure</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Martyanov, Dmitry; Soukhovitskii, Efrem; Capote, Roberto; Quesada, Jose Manuel; Chiba, Satoshi</p> <p>2017-09-01</p> <p>A new dispersive coupled-channel optical model (DCCOM) is derived that describes nucleon scattering on 238U and 232Th targets using a soft-rotator-model (SRM) description of the collective levels of the target nucleus. SRM Hamiltonian parameters are adjusted to the observed collective levels of the target nucleus. SRM nuclear wave functions (mixed in K quantum number) have been used to calculate coupling matrix elements of the generalized optical model. Five rotational <span class="hlt">bands</span> are coupled: the ground-state <span class="hlt">band</span>, β-, γ-, non-axial- <span class="hlt">bands</span>, and a negative parity <span class="hlt">band</span>. Such coupling scheme includes almost all levels below 1.2 MeV of excitation <span class="hlt">energy</span> of targets. The "effective" deformations that define inter-<span class="hlt">band</span> couplings are derived from SRM Hamiltonian parameters. Conservation of nuclear volume is enforced by introducing a monopolar deformed potential leading to <span class="hlt">additional</span> couplings between rotational <span class="hlt">bands</span>. The present DCCOM describes the total cross section differences between 238U and 232Th targets within experimental uncertainty from 50 keV up to 200 MeV of neutron incident <span class="hlt">energy</span>. SRM couplings and volume conservation allow a precise calculation of the compound-nucleus (CN) formation cross sections, which is significantly different from the one calculated with rigid-rotor potentials with any number of coupled levels.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018Nanot..29s4002A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018Nanot..29s4002A"><span>Ultraviolet photoelectron spectroscopy reveals <span class="hlt">energy-band</span> dispersion for π-stacked 7,8,15,16-tetraazaterrylene thin films in a donor–acceptor bulk heterojunction</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Aghdassi, Nabi; Wang, Qi; Ji, Ru-Ru; Wang, Bin; Fan, Jian; Duhm, Steffen</p> <p>2018-05-01</p> <p>7,8,15,16-tetraazaterrylene (TAT) thin films grown on highly oriented pyrolytic graphite (HOPG) substrates were studied extensively with regard to their intrinsic and interfacial electronic properties by means of ultraviolet photoelectron spectroscopy (UPS). Merely weak substrate–adsorbate interaction occurs at the TAT/HOPG interface, with interface energetics being only little affected by the nominal film thickness. Photon <span class="hlt">energy</span>-dependent UPS performed perpendicular to the molecular planes of TAT multilayer films at room temperature clearly reveals <span class="hlt">band</span>-like intermolecular dispersion of the TAT highest occupied molecular orbital (HOMO) <span class="hlt">energy</span>. Based on a comparison with a tight-binding model, a relatively narrow bandwidth of 54 meV is derived, which points to the presence of an intermediate regime between hopping and <span class="hlt">band</span>-like hole transport. Upon <span class="hlt">additional</span> deposition of 2,2‧:5‧,2″:5″,2″‧-quaterthiophene (4T), a 4T:TAT donor–acceptor bulk heterojunction with a considerable HOMO-level offset at the donor–acceptor interface is formed. The 4T:TAT bulk heterojunction likewise exhibits intermolecular dispersion of the TAT HOMO <span class="hlt">energy</span>, yet with a significant decreased bandwidth.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JEMat.tmp..158T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JEMat.tmp..158T"><span>Impact of Antibody Bioconjugation on Emission and <span class="hlt">Energy</span> <span class="hlt">Band</span> Profile of CdSeTe/ZnS Quantum Dots</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Torchynska, T. V.; Gomez, J. A. Jaramillo; Polupan, G.; Macotela, L. G. Vega</p> <p>2018-03-01</p> <p>The variation of the photoluminescence (PL) and Raman scattering spectra of CdSeTe/ZnS quantum dots (QDs) on conjugation to an antibody has been investigated. Two types of CdSeTe/ZnS QD with different emission wavelength (705 nm and 800 nm) were studied comparatively before and after conjugation to anti-pseudorabies virus antibody (AB). Nonconjugated QDs were characterized by Gaussian-type PL <span class="hlt">bands</span>. PL shifts to higher <span class="hlt">energy</span> and asymmetric shape of PL <span class="hlt">bands</span> was detected in PL spectra of bioconjugated QDs. The surface-enhanced Raman scattering effect was exhibited by the bioconjugated CdSeTe/ZnS QDs, indicating that the excitation light used in the Raman study generated electric dipoles in the AB molecules. The optical bandgap of the CdSeTe core was calculated numerically as a function of its radius based on an effective mass approximation model. The <span class="hlt">energy</span> <span class="hlt">band</span> diagrams for non- and bioconjugated CdSeTe/ZnS QDs were obtained, revealing a type II quantum well in the CdSeTe core. The calculations show that AB dipoles, excited in the bioconjugated QDs, stimulate a change in the <span class="hlt">energy</span> <span class="hlt">band</span> diagram of the QDs that alters the PL spectrum. These results could be useful for improving the sensitivity of QD biosensors.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27662502','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27662502"><span>High Throughput Light Absorber Discovery, Part 2: Establishing Structure-<span class="hlt">Band</span> Gap <span class="hlt">Energy</span> Relationships.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Suram, Santosh K; Newhouse, Paul F; Zhou, Lan; Van Campen, Douglas G; Mehta, Apurva; Gregoire, John M</p> <p>2016-11-14</p> <p>Combinatorial materials science strategies have accelerated materials development in a variety of fields, and we extend these strategies to enable structure-property mapping for light absorber materials, particularly in high order composition spaces. High throughput optical spectroscopy and synchrotron X-ray diffraction are combined to identify the optical properties of Bi-V-Fe oxides, leading to the identification of Bi 4 V 1.5 Fe 0.5 O 10.5 as a light absorber with direct <span class="hlt">band</span> gap near 2.7 eV. The strategic combination of experimental and data analysis techniques includes automated Tauc analysis to estimate <span class="hlt">band</span> gap <span class="hlt">energies</span> from the high throughput spectroscopy data, providing an automated platform for identifying new optical materials.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/21518425-energy-transition-characterization-mu-bands-bismuth-fiber-spectroscopy-transient-oscillations','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/21518425-energy-transition-characterization-mu-bands-bismuth-fiber-spectroscopy-transient-oscillations"><span><span class="hlt">Energy</span> transition characterization of 1.18 and 1.3 {mu}m <span class="hlt">bands</span> of bismuth fiber by spectroscopy of the transient oscillations</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Gumenyuk, Regina; Okhotnikov, Oleg G.; Golant, Konstantin</p> <p>2011-05-09</p> <p>The experimental evidence of laser transition type in bismuth-doped silica fibers operating at different spectral <span class="hlt">bands</span> is presented. Spectrally resolved transient (relaxation) oscillations studied for a Bi-doped fiber laser at room and liquid-nitrogen temperatures allow to identify the three- and four-level <span class="hlt">energy</span> <span class="hlt">bands</span>. 1.18 {mu}m short-wavelength <span class="hlt">band</span> is found to be a three-level system at room temperature with highly populated terminal <span class="hlt">energy</span> level of laser transition. The depopulation of ground level by cooling the fiber down to liquid-nitrogen temperature changes the transition to four-level type. Four-level <span class="hlt">energy</span> transition distinguished at 1.32 {mu}m exhibits the net gain at room temperature.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26459748','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26459748"><span>Electric-dipole effect of defects on the <span class="hlt">energy</span> <span class="hlt">band</span> alignment of rutile and anatase TiO₂.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zhang, Daoyu; Yang, Minnan; Dong, Shuai</p> <p>2015-11-21</p> <p>Titanium dioxide materials have been studied intensively and extensively for photocatalytic applications. A long-standing open question is the <span class="hlt">energy</span> <span class="hlt">band</span> alignment of rutile and anatase TiO2 phases, which can affect the photocatalytic process in the composite system. There are basically two contradictory viewpoints about the alignment of these two TiO2 phases supported by the respective experiments: (1) straddling type and (2) staggered type. In this work, our DFT plus U calculations show that the perfect rutile(110) and anatase(101) surfaces have the straddling type <span class="hlt">band</span> alignment, whereas the surfaces with defects can turn the <span class="hlt">band</span> alignment into the staggered type. The electric dipoles induced by defects are responsible for the reversal of <span class="hlt">band</span> alignment. Thus the defects introduced during the preparation and post-treatment processes of materials are probably the answer to the above open question regarding the <span class="hlt">band</span> alignment, which can be considered in real practice to tune the photocatalytic activity of materials.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28319838','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28319838"><span>Graphene oxide quantum dot-sensitized porous titanium dioxide microsphere: Visible-light-driven photocatalyst based on <span class="hlt">energy</span> <span class="hlt">band</span> engineering.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zhang, Yu; Qi, Fuyuan; Li, Ying; Zhou, Xin; Sun, Hongfeng; Zhang, Wei; Liu, Daliang; Song, Xi-Ming</p> <p>2017-07-15</p> <p>We report a novel graphene oxide quantum dot (GOQD)-sensitized porous TiO 2 microsphere for efficient photoelectric conversion. Electro-chemical analysis along with the Mott-Schottky equation reveals conductivity type and <span class="hlt">energy</span> <span class="hlt">band</span> structure of the two semiconductors. Based on their <span class="hlt">energy</span> <span class="hlt">band</span> structures, visible light-induced electrons can transfer from the p-type GOQD to the n-type TiO 2 . Enhanced photocurrent and photocatalytic activity in visible light further confirm the enhanced separation of electrons and holes in the nanocomposite. Copyright © 2017 Elsevier Inc. All rights reserved.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_3");'>3</a></li> <li><a href="#" onclick='return showDiv("page_4");'>4</a></li> <li class="active"><span>5</span></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_5 --> <div id="page_6" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_4");'>4</a></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li class="active"><span>6</span></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="101"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29673650','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29673650"><span>Differential recruitment of brain networks in single-digit <span class="hlt">addition</span> and multiplication: Evidence from EEG oscillations in theta and lower alpha <span class="hlt">bands</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wang, Lihan; Gan, John Q; Zhang, Li; Wang, Haixian</p> <p>2018-06-01</p> <p>Previous neuroimaging research investigating dissociation between single-digit <span class="hlt">addition</span> and multiplication has suggested that the former placed more reliance on the visuo-spatial processing whereas the latter on the verbal processing. However, there has been little exploration into the disassociation in spatio-temporal dynamics of the oscillatory brain activity in specific frequency <span class="hlt">bands</span> during the two arithmetic operations. To address this issue, the electroencephalogram (EEG) data were recorded from 19 participants engaged in a delayed verification arithmetic task. By analyzing oscillatory EEG activity in theta (5-7 Hz) and lower alpha frequency (9-10 Hz) <span class="hlt">bands</span>, we found different patterns of oscillatory brain activity between single-digit <span class="hlt">addition</span> and multiplication during the early processing stage (0-400 ms post-operand onset). Experiment results in this study showed a larger phasic increase of theta-<span class="hlt">band</span> power for <span class="hlt">addition</span> than for multiplication in the midline and the right frontal and central regions during the operator and operands presentation intervals, which was extended to the right parietal and the right occipito-temporal regions during the interval immediately after the operands presentation. In contrast, during multiplication higher phase-locking in lower alpha <span class="hlt">band</span> was evident in the centro-parietal regions during the operator presentation, which was extended to the left fronto-central and anterior regions during the operands presentation. Besides, we found stronger theta phase synchrony between the parietal areas and the right occipital areas for single-digit <span class="hlt">addition</span> than for multiplication during operands encoding. These findings of oscillatory brain activity extend the previous observations on functional dissociation between the two arithmetic operations. Copyright © 2018 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JSP...171..679M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JSP...171..679M"><span>Pair Formation of Hard Core Bosons in Flat <span class="hlt">Band</span> Systems</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mielke, Andreas</p> <p>2018-05-01</p> <p>Hard core bosons in a large class of one or two dimensional flat <span class="hlt">band</span> systems have an upper critical density, below which the ground states can be described completely. At the critical density, the ground states are Wigner crystals. If one adds a particle to the system at the critical density, the ground state and the low lying multi particle states of the system can be described as a Wigner crystal with an <span class="hlt">additional</span> pair of particles. The <span class="hlt">energy</span> <span class="hlt">band</span> for the pair is separated from the rest of the multi-particle spectrum. The proofs use a Gerschgorin type of argument for block diagonally dominant matrices. In certain one-dimensional or tree-like structures one can show that the pair is localised, for example in the chequerboard chain. For this one-dimensional system with periodic boundary condition the <span class="hlt">energy</span> <span class="hlt">band</span> for the pair is flat, the pair is localised.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22490729-edge-effects-band-gap-energy-bilayer-mos-sub-under-uniaxial-strain','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22490729-edge-effects-band-gap-energy-bilayer-mos-sub-under-uniaxial-strain"><span>Edge effects on <span class="hlt">band</span> gap <span class="hlt">energy</span> in bilayer 2H-MoS{sub 2} under uniaxial strain</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Dong, Liang; Wang, Jin; Dongare, Avinash M., E-mail: dongare@uconn.edu</p> <p>2015-06-28</p> <p>The potential of ultrathin MoS{sub 2} nanostructures for applications in electronic and optoelectronic devices requires a fundamental understanding in their electronic structure as a function of strain. Previous experimental and theoretical studies assume that an identical strain and/or stress state is always maintained in the top and bottom layers of a bilayer MoS{sub 2} film. In this study, a bilayer MoS{sub 2} supercell is constructed differently from the prototypical unit cell in order to investigate the layer-dependent electronic <span class="hlt">band</span> gap <span class="hlt">energy</span> in a bilayer MoS{sub 2} film under uniaxial mechanical deformations. The supercell contains an MoS{sub 2} bottom layer andmore » a relatively narrower top layer (nanoribbon with free edges) as a simplified model to simulate the as-grown bilayer MoS{sub 2} flakes with free edges observed experimentally. Our results show that the two layers have different <span class="hlt">band</span> gap <span class="hlt">energies</span> under a tensile uniaxial strain, although they remain mutually interacting by van der Waals interactions. The deviation in their <span class="hlt">band</span> gap <span class="hlt">energies</span> grows from 0 to 0.42 eV as the uniaxial strain increases from 0% to 6% under both uniaxial strain and stress conditions. The deviation, however, disappears if a compressive uniaxial strain is applied. These results demonstrate that tensile uniaxial strains applied to bilayer MoS{sub 2} films can result in distinct <span class="hlt">band</span> gap <span class="hlt">energies</span> in the bilayer structures. Such variations need to be accounted for when analyzing strain effects on electronic properties of bilayer or multilayered 2D materials using experimental methods or in continuum models.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1996APS..DNP..CB01F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1996APS..DNP..CB01F"><span>Identical <span class="hlt">Bands</span>: Does ``Seeing Double'' Mean We Learn Twice as Much?</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Fallon, Paul</p> <p>1996-10-01</p> <p>The phenomenon of `identical <span class="hlt">bands</span>' has been under discussion for several years, however the origin of this surprising observation, whereby rotational cascades in different nuclei exhibit very similar transition <span class="hlt">energies</span> (and/or moments of inertia), remains uncertain. The first cases of identical superdeformed <span class="hlt">bands</span> were observed when only a small number of superdeformed <span class="hlt">bands</span> were known. Since then many more examples of superdeformation have been found and it is important to see if the number of `identical' <span class="hlt">bands</span> has risen in proportion. In <span class="hlt">addition</span> the `identical' <span class="hlt">band</span> discussion has been extended to normal deformed nuclei. In this talk I will briefly review the topic of identical <span class="hlt">bands</span> and attempt to address the progress made and add some personal views on what remains to be done. Work supported in part by the U.S. DOE under contract number DE-AC03-76SF0098.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014JPhCS.557a2022W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014JPhCS.557a2022W"><span>Composite Piezoelectric Rubber <span class="hlt">Band</span> for <span class="hlt">Energy</span> Harvesting from Breathing and Limb Motion</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wang, Jhih-Jhe; Su, Huan-Jan; Hsu, Chang-I.; Su, Yu-Chuan</p> <p>2014-11-01</p> <p>We have successfully demonstrated the design and microfabrication of piezoelectric rubber <span class="hlt">bands</span> and their application in <span class="hlt">energy</span> harvesting from human motions. Composite polymeric and metallic microstructures with embedded bipolar charges are employed to realize the desired stretchability and electromechanical sensitivity. In the prototype demonstration, multilayer PDMS cellular structures coated with PTFE films and stretchable gold electrodes are fabricated and implanted with bipolar charges. The composite structures show elasticity of 300~600 kPa and extreme piezoelectricity of d33 >2000 pC/N and d31 >200 pC/N. For a working volume of 2.5cm×2.5cm×0.3mm, 10% (or 2.5mm) stretch results in effective d31 of >17000 pC/N. It is estimated that electric charge of >0.2 μC can be collected and stored per breath (or 2.5cm deformation). As such, the composite piezoelectric rubber <span class="hlt">bands</span> (with spring constants of ~200 N/m) can be mounted on elastic waistbands to harvest the circumferential stretch during breathing, or on pads around joints to harvest the elongation during limb motion. Furthermore, the wearable piezoelectric structures can be spread, stacked and connected to charge <span class="hlt">energy</span> storages and power micro devices.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1281077-theoretical-modeling-low-energy-electronic-absorption-bands-reduced-cobaloximes','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1281077-theoretical-modeling-low-energy-electronic-absorption-bands-reduced-cobaloximes"><span>Theoretical modeling of low-<span class="hlt">energy</span> electronic absorption <span class="hlt">bands</span> in reduced cobaloximes</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Bhattacharjee, Anirban; Chavarot-Kerlidou, Murielle; Dempsey, Jillian L.; ...</p> <p>2014-08-11</p> <p>Here, we report that the reduced Co(I) states of cobaloximes are powerful nucleophiles that play an important role in the hydrogen-evolving catalytic activity of these species. In this work we have analyzed the low <span class="hlt">energy</span> electronic absorption <span class="hlt">bands</span> of two cobaloxime systems experimentally and using a variety of density functional theory and molecular orbital ab initio quantum chemical approaches. Overall we find a reasonable qualitative understanding of the electronic excitation spectra of these compounds but show that obtaining quantitative results remains a challenging task.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1328964-high-throughput-light-absorber-discovery-part-establishing-structureband-gap-energy-relationships','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1328964-high-throughput-light-absorber-discovery-part-establishing-structureband-gap-energy-relationships"><span>High throughput light absorber discovery, Part 2: Establishing structure–<span class="hlt">band</span> gap <span class="hlt">energy</span> relationships</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Suram, Santosh K.; Newhouse, Paul F.; Zhou, Lan; ...</p> <p>2016-09-23</p> <p>Combinatorial materials science strategies have accelerated materials development in a variety of fields, and we extend these strategies to enable structure-property mapping for light absorber materials, particularly in high order composition spaces. High throughput optical spectroscopy and synchrotron X-ray diffraction are combined to identify the optical properties of Bi-V-Fe oxides, leading to the identification of Bi 4V 1.5Fe 0.5O 10.5 as a light absorber with direct <span class="hlt">band</span> gap near 2.7 eV. Here, the strategic combination of experimental and data analysis techniques includes automated Tauc analysis to estimate <span class="hlt">band</span> gap <span class="hlt">energies</span> from the high throughput spectroscopy data, providing an automated platformmore » for identifying new optical materials.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/1328964-high-throughput-light-absorber-discovery-part-establishing-structureband-gap-energy-relationships','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/1328964-high-throughput-light-absorber-discovery-part-establishing-structureband-gap-energy-relationships"><span>High throughput light absorber discovery, Part 2: Establishing structure–<span class="hlt">band</span> gap <span class="hlt">energy</span> relationships</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Suram, Santosh K.; Newhouse, Paul F.; Zhou, Lan</p> <p></p> <p>Combinatorial materials science strategies have accelerated materials development in a variety of fields, and we extend these strategies to enable structure-property mapping for light absorber materials, particularly in high order composition spaces. High throughput optical spectroscopy and synchrotron X-ray diffraction are combined to identify the optical properties of Bi-V-Fe oxides, leading to the identification of Bi 4V 1.5Fe 0.5O 10.5 as a light absorber with direct <span class="hlt">band</span> gap near 2.7 eV. Here, the strategic combination of experimental and data analysis techniques includes automated Tauc analysis to estimate <span class="hlt">band</span> gap <span class="hlt">energies</span> from the high throughput spectroscopy data, providing an automated platformmore » for identifying new optical materials.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JPSJ...87b4710G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JPSJ...87b4710G"><span><span class="hlt">Energy</span> <span class="hlt">Band</span> Gap Dependence of Valley Polarization of the Hexagonal Lattice</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ghalamkari, Kazu; Tatsumi, Yuki; Saito, Riichiro</p> <p>2018-02-01</p> <p>The origin of valley polarization of the hexagonal lattice is analytically discussed by tight binding method as a function of <span class="hlt">energy</span> <span class="hlt">band</span> gap. When the <span class="hlt">energy</span> gap decreases to zero, the intensity of optical absorption becomes sharp as a function of k near the K (or K') point in the hexagonal Brillouin zone, while the peak intensity at the K (or K') point keeps constant with decreasing the <span class="hlt">energy</span> gap. When the dipole vector as a function of k can have both real and imaginary parts that are perpendicular to each other in the k space, the valley polarization occurs. When the dipole vector has only real values by selecting a proper phase of wave functions, the valley polarization does not occur. The degree of the valley polarization may show a discrete change that can be relaxed to a continuous change of the degree of valley polarization when we consider the life time of photo-excited carrier.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018BrJPh..48...85S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018BrJPh..48...85S"><span>Projected Shell Model Description of Positive Parity <span class="hlt">Band</span> of 130Pr Nucleus</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Singh, Suram; Kumar, Amit; Singh, Dhanvir; Sharma, Chetan; Bharti, Arun; Bhat, G. H.; Sheikh, J. A.</p> <p>2018-02-01</p> <p>Theoretical investigation of positive parity yrast <span class="hlt">band</span> of odd-odd 130Pr nucleus is performed by applying the projected shell model. The present study is undertaken to investigate and verify the very recently observed side <span class="hlt">band</span> in 130Pr theoretically in terms of quasi-particle (qp) configuration. From the analysis of <span class="hlt">band</span> diagram, the yrast as well as side <span class="hlt">band</span> are found to arise from two-qp configuration πh 11/2 ⊗ νh 11/2. The present calculations are viewed to have qualitatively reproduced the known experimental data for yrast states, transition <span class="hlt">energies</span>, and B( M1) / B( E2) ratios of this nucleus. The recently observed positive parity side <span class="hlt">band</span> is also reproduced by the present calculations. The <span class="hlt">energy</span> states of the side <span class="hlt">band</span> are predicted up to spin 25+, which is far above the known experimental spin of 18+ and this could serve as a motivational factor for future experiments. In <span class="hlt">addition</span>, the reduced transition probability B( E2) for interband transitions has also been calculated for the first time in projected shell model, which would serve as an encouragement for other research groups in the future.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2005APS..MARV14002E','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2005APS..MARV14002E"><span>Interface <span class="hlt">band</span> alignment in high-k gate stacks</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Eric, Bersch; Hartlieb, P.</p> <p>2005-03-01</p> <p>In order to successfully implement alternate high-K dielectric materials into MOS structures, the interface properties of MOS gate stacks must be better understood. Dipoles that may form at the metal/dielectric and dielectric/semiconductor interfaces make the <span class="hlt">band</span> offsets difficult to predict. We have measured the conduction and valence <span class="hlt">band</span> densities of states for a variety MOS stacks using in situ using inverse photoemission (IPE) and photoemission spectroscopy (PES), respectively. Results obtained from clean and metallized (with Ru or Al) HfO2/Si, SiO2/Si and mixed silicate films will be presented. IPE indicates a shift of the conduction <span class="hlt">band</span> minimum (CBM) to higher <span class="hlt">energy</span> (i.e. away from EF) with increasing SiO2. The effect of metallization on the location of <span class="hlt">band</span> edges depends upon the metal species. The <span class="hlt">addition</span> of N to the dielectrics shifts the CBM in a way that is thickness dependent. Possible mechanisms for these observed effects will be discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.dtic.mil/docs/citations/ADA455492','DTIC-ST'); return false;" href="http://www.dtic.mil/docs/citations/ADA455492"><span>Enhanced Spontaneous Emission of Bloch Oscillation Radiation from a Single <span class="hlt">Energy</span> <span class="hlt">Band</span></span></a></p> <p><a target="_blank" href="http://www.dtic.mil/">DTIC Science & Technology</a></p> <p></p> <p>2006-06-30</p> <p>ignore interband tunneling , spon- taneous photon emission occurs as the Bloch electron inter- acts with the quantum radiation field; the emission occurs... interband coupling 17 and electron intraband scattering are ignored. Therefore, the quantum dynamics is described by the time-dependent Schrödinger...single <span class="hlt">band</span> “n0” of a periodic crystal with <span class="hlt">energy</span> n0K; the ef- fects of interband coupling15 and electron intraband scatter- ing are ignored</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JSV...428..119R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JSV...428..119R"><span>Cross-frequency and <span class="hlt">band</span>-averaged response variance prediction in the hybrid deterministic-statistical <span class="hlt">energy</span> analysis method</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Reynders, Edwin P. B.; Langley, Robin S.</p> <p>2018-08-01</p> <p>The hybrid deterministic-statistical <span class="hlt">energy</span> analysis method has proven to be a versatile framework for modeling built-up vibro-acoustic systems. The stiff system components are modeled deterministically, e.g., using the finite element method, while the wave fields in the flexible components are modeled as diffuse. In the present paper, the hybrid method is extended such that not only the ensemble mean and variance of the harmonic system response can be computed, but also of the <span class="hlt">band</span>-averaged system response. This variance represents the uncertainty that is due to the assumption of a diffuse field in the flexible components of the hybrid system. The developments start with a cross-frequency generalization of the reciprocity relationship between the total <span class="hlt">energy</span> in a diffuse field and the cross spectrum of the blocked reverberant loading at the boundaries of that field. By making extensive use of this generalization in a first-order perturbation analysis, explicit expressions are derived for the cross-frequency and <span class="hlt">band</span>-averaged variance of the vibrational <span class="hlt">energies</span> in the diffuse components and for the cross-frequency and <span class="hlt">band</span>-averaged variance of the cross spectrum of the vibro-acoustic field response of the deterministic components. These expressions are extensively validated against detailed Monte Carlo analyses of coupled plate systems in which diffuse fields are simulated by randomly distributing small point masses across the flexible components, and good agreement is found.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AIPA....7f5005K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AIPA....7f5005K"><span>Local strain-induced <span class="hlt">band</span> gap fluctuations and exciton localization in aged WS2 monolayers</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Krustok, J.; Kaupmees, R.; Jaaniso, R.; Kiisk, V.; Sildos, I.; Li, B.; Gong, Y.</p> <p>2017-06-01</p> <p>Optical properties of aged WS2 monolayers grown by CVD method on Si/SiO2 substrates are studied using temperature dependent photoluminescence and reflectance contrast spectroscopy. Aged WS2 monolayers have a typical surface roughness about 0.5 nm and, in <span class="hlt">addition</span>, a high density of nanoparticles (nanocaps) with the base diameter about 30 nm and average height of 7 nm. The A-exciton of aged monolayer has a peak position at 1.951 eV while in as-grown monolayer the peak is at about 24 meV higher <span class="hlt">energy</span> at room temperature. This red-shift is explained using local tensile strain concept, where strain value of 2.1% was calculated for these nanocap regions. Strained nanocaps have lower <span class="hlt">band</span> gap <span class="hlt">energy</span> and excitons will funnel into these regions. At T=10K a double exciton and trion peaks were revealed. The separation between double peaks is about 20 meV and the origin of higher <span class="hlt">energy</span> peaks is related to the optical <span class="hlt">band</span> gap <span class="hlt">energy</span> fluctuations caused by random distribution of local tensile strain due to increased surface roughness. In <span class="hlt">addition</span>, a wide defect related exciton <span class="hlt">band</span> XD was found at about 1.93 eV in all aged monolayers. It is shown that the theory of localized excitons describes well the temperature dependence of peak position and halfwidth of the A-exciton <span class="hlt">band</span>. The possible origin of nanocaps is also discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AIPC.1901b0014P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AIPC.1901b0014P"><span>Application of back-propagation artificial neural network (ANN) to predict crystallite size and <span class="hlt">band</span> gap <span class="hlt">energy</span> of ZnO quantum dots</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Pelicano, Christian Mark; Rapadas, Nick; Cagatan, Gerard; Magdaluyo, Eduardo</p> <p>2017-12-01</p> <p>Herein, the crystallite size and <span class="hlt">band</span> gap <span class="hlt">energy</span> of zinc oxide (ZnO) quantum dots were predicted using artificial neural network (ANN). Three input factors including reagent ratio, growth time, and growth temperature were examined with respect to crystallite size and <span class="hlt">band</span> gap <span class="hlt">energy</span> as response factors. The generated results from neural network model were then compared with the experimental results. Experimental crystallite size and <span class="hlt">band</span> gap <span class="hlt">energy</span> of ZnO quantum dots were measured from TEM images and absorbance spectra, respectively. The Levenberg-Marquardt (LM) algorithm was used as the learning algorithm for the ANN model. The performance of the ANN model was then assessed through mean square error (MSE) and regression values. Based on the results, the ANN modelling results are in good agreement with the experimental data.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017APS..MARF33005Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017APS..MARF33005Z"><span>Symmetries and <span class="hlt">band</span> gaps in nanoribbons</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zhang, Zhiwei; Tian, Yiteng; Fernando, Gayanath; Kocharian, Armen</p> <p></p> <p>In ideal graphene-like systems, time reversal and sublattice symmetries preserve the degeneracies at the Dirac point(s). We have examined such degeneracies in the <span class="hlt">band</span> structure as well as the transport properties in various arm-twisted (graphene-related) nanoribbons. A twist angle is defined such that at 0 degrees the ribbon is a rectangular ribbon and at 60 degrees the ribbon is cut from a honeycomb lattice. Using model Hamiltonians and first principles calculations in these nanoribbons with Z2 topology, we have monitored the <span class="hlt">band</span> structure as a function of the twist angle θ. In twisted ribbons, it turns out that the introduction of an extra hopping term leads to a gap opening. We have also calculated the size and temperature broadening effects in similar ribbons in <span class="hlt">addition</span> to Rashba-induced transport properties. The authors acknowledge the computing facilities provided by the Center for Functional Nanomaterials, Brookhaven National Laboratory supported by the U.S. Department of <span class="hlt">Energy</span>, Office of Basic <span class="hlt">Energy</span> Sciences, under Contract No.DE-AC02- 98CH10886.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23350611','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23350611"><span>Fused dithienogermolodithiophene low <span class="hlt">band</span> gap polymers for high-performance organic solar cells without processing <span class="hlt">additives</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zhong, Hongliang; Li, Zhe; Deledalle, Florent; Fregoso, Elisa Collado; Shahid, Munazza; Fei, Zhuping; Nielsen, Christian B; Yaacobi-Gross, Nir; Rossbauer, Stephan; Anthopoulos, Thomas D; Durrant, James R; Heeney, Martin</p> <p>2013-02-13</p> <p>We report the synthesis of a novel ladder-type fused ring donor, dithienogermolodithiophene, in which two thieno[3,2-b]thiophene units are held coplanar by a bridging dialkyl germanium. Polymerization of this extended monomer with N-octylthienopyrrolodione by Stille polycondensation afforded a polymer, pDTTG-TPD, with an optical <span class="hlt">band</span> gap of 1.75 eV combined with a high ionization potential. Bulk heterojunction solar cells based upon pDTTG-TPD:PC(71)BM blends afforded efficiencies up to 7.2% without the need for thermal annealing or processing <span class="hlt">additives</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28743864','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28743864"><span>Optical <span class="hlt">Band</span> Gap Alteration of Graphene Oxide via Ozone Treatment.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Hasan, Md Tanvir; Senger, Brian J; Ryan, Conor; Culp, Marais; Gonzalez-Rodriguez, Roberto; Coffer, Jeffery L; Naumov, Anton V</p> <p>2017-07-25</p> <p>Graphene oxide (GO) is a graphene derivative that emits fluorescence, which makes GO an attractive material for optoelectronics and biotechnology. In this work, we utilize ozone treatment to controllably tune the <span class="hlt">band</span> gap of GO, which can significantly enhance its applications. Ozone treatment in aqueous GO suspensions yields the <span class="hlt">addition</span>/rearrangement of oxygen-containing functional groups suggested by the increase in vibrational transitions of C-O and C=O moieties. Concomitantly it leads to an initial increase in GO fluorescence intensity and significant (100 nm) blue shifts in emission maxima. Based on the model of GO fluorescence originating from sp 2 graphitic islands confined by oxygenated addends, we propose that ozone-induced functionalization decreases the size of graphitic islands affecting the GO <span class="hlt">band</span> gap and emission <span class="hlt">energies</span>. TEM analyses of GO flakes confirm the size decrease of ordered sp 2 domains with ozone treatment, whereas semi-empirical PM3 calculations on model addend-confined graphitic clusters predict the inverse dependence of the <span class="hlt">band</span> gap <span class="hlt">energies</span> on sp 2 cluster size. This model explains ozone-induced increase in emission <span class="hlt">energies</span> yielding fluorescence blue shifts and helps develop an understanding of the origins of GO fluorescence emission. Furthermore, ozone treatment provides a versatile approach to controllably alter GO <span class="hlt">band</span> gap for optoelectronics and bio-sensing applications.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1987ZPhyA.328..399Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1987ZPhyA.328..399Z"><span>In-beam spectroscopy of the k π=0- <span class="hlt">bands</span> in230 236U</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zeyen, P.; Ackermann, B.; Dämmrich, U.; Euler, K.; Grafen, V.; Günther, C.; Herzog, P.; Marten-Tölle, M.; Prillwitz, B.; Tölle, R.; Lauterbach, Ch.; Maier, H. J.</p> <p>1987-12-01</p> <p>The K π=0- <span class="hlt">bands</span> in even uranium nuclei were studied in the compound reactions231Pa( p, 2 n)230U,230, 232Th( α,2 n)232, 234U and236U( d, pn)236U. In-beam γ-rays were measured in coincidence with conversion-electrons, which were detected with an iron-free orange spectrometer. The negative-parity levels are observed up to intermediate spins ( I<13-). In <span class="hlt">addition</span>, the 1- and 3- levels in230U were confirmed by a decay study with an isotope separated230Pa source. For the heavier isotopes ( A≥232) the properties of the K π=0- <span class="hlt">bands</span> (<span class="hlt">energies</span> and γ-branchings) are consistent with a vibrational character of these <span class="hlt">bands</span>. For230U the K π=0- <span class="hlt">band</span> lies at rather low <span class="hlt">energy</span> ( E(1-)=367 keV), and the level spacings within this <span class="hlt">band</span> are very similar to those of the isotones228Th and226Ra, which might indicate the onset of a stable octupole deformation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22486039-elimination-surface-band-bending-polar-inn-thin-gan-capping','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22486039-elimination-surface-band-bending-polar-inn-thin-gan-capping"><span>Elimination of surface <span class="hlt">band</span> bending on N-polar InN with thin GaN capping</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Kuzmík, J., E-mail: Jan.Kuzmik@savba.sk; Haščík, Š.; Kučera, M.</p> <p>2015-11-09</p> <p>0.5–1 μm thick InN (0001) films grown by molecular-beam epitaxy with N- or In-polarity are investigated for the presence of native oxide, surface <span class="hlt">energy</span> <span class="hlt">band</span> bending, and effects introduced by 2 to 4 monolayers of GaN capping. Ex situ angle-resolved x-ray photo-electron spectroscopy is used to construct near-surface (GaN)/InN <span class="hlt">energy</span> profiles, which is combined with deconvolution of In3d signal to trace the presence of InN native oxide for different types of polarity and capping. Downwards surface <span class="hlt">energy</span> <span class="hlt">band</span> bending was observed on bare samples with native oxide, regardless of the polarity. It was found that the In-polar InN surface is mostmore » readily oxidized, however, with only slightly less <span class="hlt">band</span> bending if compared with the N-polar sample. On the other hand, InN surface oxidation was effectively mitigated by GaN capping. Still, as confirmed by ultra-violet photo-electron spectroscopy and by <span class="hlt">energy</span> <span class="hlt">band</span> diagram calculations, thin GaN cap layer may provide negative piezoelectric polarization charge at the GaN/InN hetero-interface of the N-polar sample, in <span class="hlt">addition</span> to the passivation effect. These effects raised the <span class="hlt">band</span> diagram up by about 0.65 eV, reaching a flat-<span class="hlt">band</span> profile.« less</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_4");'>4</a></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li class="active"><span>6</span></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_6 --> <div id="page_7" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li class="active"><span>7</span></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="121"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016OptMa..58...51C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016OptMa..58...51C"><span>Optical <span class="hlt">band</span> gaps of organic semiconductor materials</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Costa, José C. S.; Taveira, Ricardo J. S.; Lima, Carlos F. R. A. C.; Mendes, Adélio; Santos, Luís M. N. B. F.</p> <p>2016-08-01</p> <p>UV-Vis can be used as an easy and forthright technique to accurately estimate the <span class="hlt">band</span> gap <span class="hlt">energy</span> of organic π-conjugated materials, widely used as thin films/composites in organic and hybrid electronic devices such as OLEDs, OPVs and OFETs. The electronic and optical properties, including HOMO-LUMO <span class="hlt">energy</span> gaps of π-conjugated systems were evaluated by UV-Vis spectroscopy in CHCl3 solution for a large number of relevant π-conjugated systems: tris-8-hydroxyquinolinatos (Alq3, Gaq3, Inq3, Al(qNO2)3, Al(qCl)3, Al(qBr)3, In(qNO2)3, In(qCl)3 and In(qBr)3); triphenylamine derivatives (DDP, p-TTP, TPB, TPD, TDAB, m-MTDAB, NPB, α-NPD); oligoacenes (naphthalene, anthracene, tetracene and rubrene); oligothiophenes (α-2T, β-2T, α-3T, β-3T, α-4T and α-5T). <span class="hlt">Additionally</span>, some electronic properties were also explored by quantum chemical calculations. The experimental UV-Vis data are in accordance with the DFT predictions and indicate that the <span class="hlt">band</span> gap <span class="hlt">energies</span> of the OSCs dissolved in CHCl3 solution are consistent with the values presented for thin films.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018AcAau.142..221A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AcAau.142..221A"><span>Concentrated <span class="hlt">energy</span> <span class="hlt">addition</span> for active drag reduction in hypersonic flow regime</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ashwin Ganesh, M.; John, Bibin</p> <p>2018-01-01</p> <p>Numerical optimization of hypersonic drag reduction technique based on concentrated <span class="hlt">energy</span> <span class="hlt">addition</span> is presented in this study. A reduction in wave drag is realized through concentrated <span class="hlt">energy</span> <span class="hlt">addition</span> in the hypersonic flowfield upstream of the blunt body. For the exhaustive optimization presented in this study, an in-house high precision inviscid flow solver has been developed. Studies focused on the identification of "optimum <span class="hlt">energy</span> <span class="hlt">addition</span> location" have revealed the existence of multiple minimum drag points. The wave drag coefficient is observed to drop from 0.85 to 0.45 when 50 Watts of <span class="hlt">energy</span> is added to an <span class="hlt">energy</span> bubble of 1 mm radius located at 74.7 mm upstream of the stagnation point. A direct proportionality has been identified between <span class="hlt">energy</span> bubble size and wave drag coefficient. Dependence of drag coefficient on the upstream added <span class="hlt">energy</span> magnitude is also revealed. Of the observed multiple minimum drag points, the <span class="hlt">energy</span> deposition point (EDP) that offers minimum wave drag just after a sharp drop in drag is proposed as the most optimum <span class="hlt">energy</span> <span class="hlt">addition</span> location.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29393900','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29393900"><span>Accurate <span class="hlt">Energy</span> Consumption Modeling of IEEE 802.15.4e TSCH Using Dual-<span class="hlt">Band</span>OpenMote Hardware.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Daneels, Glenn; Municio, Esteban; Van de Velde, Bruno; Ergeerts, Glenn; Weyn, Maarten; Latré, Steven; Famaey, Jeroen</p> <p>2018-02-02</p> <p>The Time-Slotted Channel Hopping (TSCH) mode of the IEEE 802.15.4e amendment aims to improve reliability and <span class="hlt">energy</span> efficiency in industrial and other challenging Internet-of-Things (IoT) environments. This paper presents an accurate and up-to-date <span class="hlt">energy</span> consumption model for devices using this IEEE 802.15.4e TSCH mode. The model identifies all network-related CPU and radio state changes, thus providing a precise representation of the device behavior and an accurate prediction of its <span class="hlt">energy</span> consumption. Moreover, <span class="hlt">energy</span> measurements were performed with a dual-<span class="hlt">band</span> OpenMote device, running the OpenWSN firmware. This allows the model to be used for devices using 2.4 GHz, as well as 868 MHz. Using these measurements, several network simulations were conducted to observe the TSCH <span class="hlt">energy</span> consumption effects in end-to-end communication for both frequency <span class="hlt">bands</span>. Experimental verification of the model shows that it accurately models the consumption for all possible packet sizes and that the calculated consumption on average differs less than 3% from the measured consumption. This deviation includes measurement inaccuracies and the variations of the guard time. As such, the proposed model is very suitable for accurate <span class="hlt">energy</span> consumption modeling of TSCH networks.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5855993','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5855993"><span>Accurate <span class="hlt">Energy</span> Consumption Modeling of IEEE 802.15.4e TSCH Using Dual-<span class="hlt">Band</span>OpenMote Hardware</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Municio, Esteban; Van de Velde, Bruno; Latré, Steven</p> <p>2018-01-01</p> <p>The Time-Slotted Channel Hopping (TSCH) mode of the IEEE 802.15.4e amendment aims to improve reliability and <span class="hlt">energy</span> efficiency in industrial and other challenging Internet-of-Things (IoT) environments. This paper presents an accurate and up-to-date <span class="hlt">energy</span> consumption model for devices using this IEEE 802.15.4e TSCH mode. The model identifies all network-related CPU and radio state changes, thus providing a precise representation of the device behavior and an accurate prediction of its <span class="hlt">energy</span> consumption. Moreover, <span class="hlt">energy</span> measurements were performed with a dual-<span class="hlt">band</span> OpenMote device, running the OpenWSN firmware. This allows the model to be used for devices using 2.4 GHz, as well as 868 MHz. Using these measurements, several network simulations were conducted to observe the TSCH <span class="hlt">energy</span> consumption effects in end-to-end communication for both frequency <span class="hlt">bands</span>. Experimental verification of the model shows that it accurately models the consumption for all possible packet sizes and that the calculated consumption on average differs less than 3% from the measured consumption. This deviation includes measurement inaccuracies and the variations of the guard time. As such, the proposed model is very suitable for accurate <span class="hlt">energy</span> consumption modeling of TSCH networks. PMID:29393900</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1997APS..PAC..7P48N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1997APS..PAC..7P48N"><span>High Peak Power Test and Evaluation of S-<span class="hlt">band</span> Waveguide Switches</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Nassiri, A.; Grelick, A.; Kustom, R. L.; White, M.</p> <p>1997-05-01</p> <p>The injector and source of particles for the Advanced Photon Source is a 2856-MHz S-<span class="hlt">band</span> electron-positron linear accelerator (linac) which produces electrons with <span class="hlt">energies</span> up to 650 MeV or positrons with <span class="hlt">energies</span> up to 450 MeV. To improve the linac rf system availability, an <span class="hlt">additional</span> modulator-klystron subsystem is being constructed to provide a switchable hot spare unit for each of the five exsisting S-<span class="hlt">band</span> transmitters. The switching of the transmitters will require the use of SF6-pressurized S-<span class="hlt">band</span> waveguide switches at a peak operating power of 35 MW. Such rf switches have been successfully operated at other accelerator facilities but at lower peak powers. A test stand has been set up at the Stanford Linear Accelerator Center (SLAC) Klystron Factory to conduct tests comparing the power handling characteristics of two WR-284 and one WR-340 switches. Test results are presented and their implications for the design of the switching system are discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22261629-camel-back-band-induced-power-factor-enhancement-thermoelectric-lead-tellurium-from-boltzmann-transport-calculations','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22261629-camel-back-band-induced-power-factor-enhancement-thermoelectric-lead-tellurium-from-boltzmann-transport-calculations"><span>Camel-back <span class="hlt">band</span>-induced power factor enhancement of thermoelectric lead-tellurium from Boltzmann transport calculations</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Wang, X. G., E-mail: wang2006@mail.ustc.edu.cn; Wang, L., E-mail: sqtb@mail.ustc.edu.cn; Liu, J., E-mail: jingliu@mail.ustc.edu.cn</p> <p>2014-03-31</p> <p><span class="hlt">Band</span> structures of PbTe can be abnormally bended via dual-doping on both the cationic and anionic sites to form camel-back multivalley <span class="hlt">energy</span> <span class="hlt">band</span> structures near the <span class="hlt">band</span> edge. As a result, <span class="hlt">additional</span> carrier pockets and strong intervalley scattering of carriers are introduced. Boltzmann transport calculations indicate that their contradictory effects yield remarkably enhanced power factor due to the improved thermopower and almost unchanged electrical conductivity in low temperature and high carrier concentration ranges. These findings prove dual-doping-induced <span class="hlt">band</span> bending as an effective approach to improve the thermoelectric properties of PbTe and other similar materials.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016EL....11448001S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016EL....11448001S"><span><span class="hlt">Energy</span> <span class="hlt">band</span> gaps in graphene nanoribbons with corners</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Szczȩśniak, Dominik; Durajski, Artur P.; Khater, Antoine; Ghader, Doried</p> <p>2016-05-01</p> <p>In the present paper, we study the relation between the <span class="hlt">band</span> gap size and the corner-corner length in representative chevron-shaped graphene nanoribbons (CGNRs) with 120° and 150° corner edges. The direct physical insight into the electronic properties of CGNRs is provided within the tight-binding model with phenomenological edge parameters, developed against recent first-principle results. We show that the analyzed CGNRs exhibit inverse relation between their <span class="hlt">band</span> gaps and corner-corner lengths, and that they do not present a metal-insulator transition when the chemical edge modifications are introduced. Our results also suggest that the <span class="hlt">band</span> gap width for the CGNRs is predominantly governed by the armchair edge effects, and is tunable through edge modifications with foreign atoms dressing.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19920037746&hterms=Einstein&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3DEinstein','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19920037746&hterms=Einstein&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3DEinstein"><span>Einstein coefficients for rotational lines of the (0,0) <span class="hlt">band</span> of the NO A2sigma(+)-X2Pi system</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Reisel, John R.; Carter, Campbell D.; Laurendeau, Normand M.</p> <p>1992-01-01</p> <p>A summary of the spectroscopic equations necessary for prediction of the molecular transition <span class="hlt">energies</span> and the Einstein A and B coefficients for rovibronic lines of the gamma(0,0) <span class="hlt">band</span> of nitric oxide (NO) is presented. The calculated molecular transition <span class="hlt">energies</span> are all within 0.57/cm of published experimental values; in <span class="hlt">addition</span>, over 95 percent of the calculated <span class="hlt">energies</span> give agreement with measured results within 0.25/cm. Einstein coefficients are calculated from the <span class="hlt">band</span> A00 value and the known Hoenl-London factors and are tabulated for individual rovibronic transitions in the NO A2sigma(+)-X2Pi(0,0) <span class="hlt">band</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015ApPhL.107f2104K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015ApPhL.107f2104K"><span>Compositional bowing of <span class="hlt">band</span> <span class="hlt">energies</span> and their deformation potentials in strained InGaAs ternary alloys: A first-principles study</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Khomyakov, Petr A.; Luisier, Mathieu; Schenk, Andreas</p> <p>2015-08-01</p> <p>Using first-principles calculations, we show that the conduction and valence <span class="hlt">band</span> <span class="hlt">energies</span> and their deformation potentials exhibit a non-negligible compositional bowing in strained ternary semiconductor alloys such as InGaAs. The electronic structure of these compounds has been calculated within the framework of local density approximation and hybrid functional approach for large cubic supercells and special quasi-random structures, which represent two kinds of model structures for random alloys. We find that the predicted bowing effect for the <span class="hlt">band</span> <span class="hlt">energy</span> deformation potentials is rather insensitive to the choice of the functional and alloy structural model. The direction of bowing is determined by In cations that give a stronger contribution to the formation of the InxGa1-xAs valence <span class="hlt">band</span> states with x ≳ 0.5, compared to Ga cations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1981SSCom..39..831K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1981SSCom..39..831K"><span>Positron and electron <span class="hlt">energy</span> <span class="hlt">bands</span> in several ionic crystals using restricted Hartree-Fock method</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kunz, A. B.; Waber, J. T.</p> <p>1981-08-01</p> <p>Using a restricted Hartree-Fock formalism and suitably localized and symmetrized wave functions, both the positron and electron <span class="hlt">energy</span> <span class="hlt">bands</span> were calculated for NaF, MgO and NiO. The lowest positron state at Γ 1 lies above the vacuum level and negative work functions are predicted. Positron annihilation rates were calculated and found to be in good agreement with measured lifetimes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20150003431','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20150003431"><span>Single-<span class="hlt">Band</span> and Dual-<span class="hlt">Band</span> Infrared Detectors</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Ting, David Z. (Inventor); Gunapala, Sarath D. (Inventor); Soibel, Alexander (Inventor); Nguyen, Jean (Inventor); Khoshakhlagh, Arezou (Inventor)</p> <p>2015-01-01</p> <p>Bias-switchable dual-<span class="hlt">band</span> infrared detectors and methods of manufacturing such detectors are provided. The infrared detectors are based on a back-to-back heterojunction diode design, where the detector structure consists of, sequentially, a top contact layer, a unipolar hole barrier layer, an absorber layer, a unipolar electron barrier, a second absorber, a second unipolar hole barrier, and a bottom contact layer. In <span class="hlt">addition</span>, by substantially reducing the width of one of the absorber layers, a single-<span class="hlt">band</span> infrared detector can also be formed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20170004927','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20170004927"><span>Single-<span class="hlt">Band</span> and Dual-<span class="hlt">Band</span> Infrared Detectors</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Ting, David Z. (Inventor); Gunapala, Sarath D. (Inventor); Soibel, Alexander (Inventor); Nguyen, Jean (Inventor); Khoshakhlagh, Arezou (Inventor)</p> <p>2017-01-01</p> <p>Bias-switchable dual-<span class="hlt">band</span> infrared detectors and methods of manufacturing such detectors are provided. The infrared detectors are based on a back-to-back heterojunction diode design, where the detector structure consists of, sequentially, a top contact layer, a unipolar hole barrier layer, an absorber layer, a unipolar electron barrier, a second absorber, a second unipolar hole barrier, and a bottom contact layer. In <span class="hlt">addition</span>, by substantially reducing the width of one of the absorber layers, a single-<span class="hlt">band</span> infrared detector can also be formed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22412964-energy-band-alignment-electronic-states-amorphous-carbon-surfaces-vacuo-aqueous-environment','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22412964-energy-band-alignment-electronic-states-amorphous-carbon-surfaces-vacuo-aqueous-environment"><span><span class="hlt">Energy</span> <span class="hlt">band</span> alignment and electronic states of amorphous carbon surfaces in vacuo and in aqueous environment</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Caro, Miguel A., E-mail: mcaroba@gmail.com; Department of Applied Physics, COMP Centre of Excellence in Computational Nanoscience, Aalto University, Espoo; Määttä, Jukka</p> <p>2015-01-21</p> <p>In this paper, we obtain the <span class="hlt">energy</span> <span class="hlt">band</span> positions of amorphous carbon (a–C) surfaces in vacuum and in aqueous environment. The calculations are performed using a combination of (i) classical molecular dynamics (MD), (ii) Kohn-Sham density functional theory with the Perdew-Burke-Ernzerhof (PBE) exchange-correlation functional, and (iii) the screened-exchange hybrid functional of Heyd, Scuseria, and Ernzerhof (HSE). PBE allows an accurate generation of a-C and the evaluation of the local electrostatic potential in the a-C/water system, HSE yields an improved description of energetic positions which is critical in this case, and classical MD enables a computationally affordable description of water. Ourmore » explicit calculation shows that, both in vacuo and in aqueous environment, the a-C electronic states available in the region comprised between the H{sub 2}/H{sub 2}O and O{sub 2}/H{sub 2}O levels of water correspond to both occupied and unoccupied states within the a-C pseudogap region. These are localized states associated to sp{sup 2} sites in a-C. The <span class="hlt">band</span> realignment induces a shift of approximately 300 meV of the a-C <span class="hlt">energy</span> <span class="hlt">band</span> positions with respect to the redox levels of water.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22416163-non-pairwise-additivity-leading-order-dispersion-energy','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22416163-non-pairwise-additivity-leading-order-dispersion-energy"><span>Non-pairwise <span class="hlt">additivity</span> of the leading-order dispersion <span class="hlt">energy</span></span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Hollett, Joshua W., E-mail: j.hollett@uwinnipeg.ca</p> <p>2015-02-28</p> <p>The leading-order (i.e., dipole-dipole) dispersion <span class="hlt">energy</span> is calculated for one-dimensional (1D) and two-dimensional (2D) infinite lattices, and an infinite 1D array of infinitely long lines, of doubly occupied locally harmonic wells. The dispersion <span class="hlt">energy</span> is decomposed into pairwise and non-pairwise <span class="hlt">additive</span> components. By varying the force constant and separation of the wells, the non-pairwise <span class="hlt">additive</span> contribution to the dispersion <span class="hlt">energy</span> is shown to depend on the overlap of density between neighboring wells. As well separation is increased, the non-pairwise <span class="hlt">additivity</span> of the dispersion <span class="hlt">energy</span> decays. The different rates of decay for 1D and 2D lattices of wells is explained inmore » terms of a Jacobian effect that influences the number of nearest neighbors. For an array of infinitely long lines of wells spaced 5 bohrs apart, and an inter-well spacing of 3 bohrs within a line, the non-pairwise <span class="hlt">additive</span> component of the leading-order dispersion <span class="hlt">energy</span> is −0.11 kJ mol{sup −1} well{sup −1}, which is 7% of the total. The polarizability of the wells and the density overlap between them are small in comparison to that of the atomic densities that arise from the molecular density partitioning used in post-density-functional theory (DFT) damped dispersion corrections, or DFT-D methods. Therefore, the nonadditivity of the leading-order dispersion observed here is a conservative estimate of that in molecular clusters.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018PhyE...97..401T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PhyE...97..401T"><span>Anomalies in the 1D Anderson model: Beyond the <span class="hlt">band</span>-centre and <span class="hlt">band</span>-edge cases</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Tessieri, L.; Izrailev, F. M.</p> <p>2018-03-01</p> <p>We consider the one-dimensional Anderson model with weak disorder. Using the Hamiltonian map approach, we analyse the validity of the random-phase approximation for resonant values of the <span class="hlt">energy</span>, E = 2 cos(πr) , with r a rational number. We expand the invariant measure of the phase variable in powers of the disorder strength and we show that, contrary to what happens at the centre and at the edges of the <span class="hlt">band</span>, for all other resonant <span class="hlt">energies</span> the leading term of the invariant measure is uniform. When higher-order terms are taken into account, a modulation of the invariant measure appears for all resonant values of the <span class="hlt">energy</span>. This implies that, when the localisation length is computed within the second-order approximation in the disorder strength, the Thouless formula is valid everywhere except at the <span class="hlt">band</span> centre and at the <span class="hlt">band</span> edges.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28779768','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28779768"><span><span class="hlt">Energy</span>-effective Grinding of Inorganic Solids Using Organic <span class="hlt">Additives</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Mishra, Ratan K; Weibel, Martin; Müller, Thomas; Heinz, Hendrik; Flatt, Robert J</p> <p>2017-08-09</p> <p>We present our research findings related to new formulations of the organic <span class="hlt">additives</span> (grinding aids) needed for the efficient grinding of inorganic solids. Even though the size reduction phenomena of the inorganic solid particles in a ball mill is purely a physical process, the <span class="hlt">addition</span> of grinding aids in milling media introduces a complex physicochemical process. In <span class="hlt">addition</span> to further gain in productivity, the organic <span class="hlt">additive</span> helps to reduce the <span class="hlt">energy</span> needed for grinding, which in the case of cement clinker has major environmental implications worldwide. This is primarily due to the tremendous amounts of cement produced and almost 30% of the associated electrical <span class="hlt">energy</span> is consumed for grinding. In this paper, we examine the question of how to optimize these grinding aids linking molecular insight into their working mechanisms, and also how to design chemical <span class="hlt">additives</span> of improved performance for industrial comminution.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JQSRT.210..127P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JQSRT.210..127P"><span>Potential <span class="hlt">energy</span> surface, dipole moment surface and the intensity calculations for the 10 μm, 5 μm and 3 μm <span class="hlt">bands</span> of ozone</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Polyansky, Oleg L.; Zobov, Nikolai F.; Mizus, Irina I.; Kyuberis, Aleksandra A.; Lodi, Lorenzo; Tennyson, Jonathan</p> <p>2018-05-01</p> <p>Monitoring ozone concentrations in the Earth's atmosphere using spectroscopic methods is a major activity which undertaken both from the ground and from space. However there are long-running issues of consistency between measurements made at infrared (IR) and ultraviolet (UV) wavelengths. In <span class="hlt">addition</span>, key O3 IR <span class="hlt">bands</span> at 10 μm, 5 μm and 3 μm also yield results which differ by a few percent when used for retrievals. These problems stem from the underlying laboratory measurements of the line intensities. Here we use quantum chemical techniques, first principles electronic structure and variational nuclear-motion calculations, to address this problem. A new high-accuracy ab initio dipole moment surface (DMS) is computed. Several spectroscopically-determined potential <span class="hlt">energy</span> surfaces (PESs) are constructed by fitting to empirical <span class="hlt">energy</span> levels in the region below 7000 cm-1 starting from an ab initio PES. Nuclear motion calculations using these new surfaces allow the unambiguous determination of the intensities of 10 μm <span class="hlt">band</span> transitions, and the computation of the intensities of 10 μm and 5 μm <span class="hlt">bands</span> within their experimental error. A decrease in intensities within the 3 μm is predicted which appears consistent with atmospheric retrievals. The PES and DMS form a suitable starting point both for the computation of comprehensive ozone line lists and for future calculations of electronic transition intensities.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4294628','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4294628"><span>A Fluorescent Indicator for Imaging Lysosomal Zinc(II) with Förster Resonance <span class="hlt">Energy</span> Transfer (FRET)-Enhanced Photostability and a Narrow <span class="hlt">Band</span> of Emission</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Sreenath, Kesavapillai; Yuan, Zhao; Allen, John R.</p> <p>2015-01-01</p> <p>We demonstrate a strategy to transfer the zinc(II) sensitivity of a fluoroionophore with low photostability and a broad emission <span class="hlt">band</span> to a bright and photostable fluorophore with a narrow emission <span class="hlt">band</span>. The two fluorophores are covalently connected to afford an intramolecular Förster resonance <span class="hlt">energy</span> transfer (FRET) conjugate. The FRET donor in the conjugate is a zinc(II)-sensitive arylvinylbipyridyl fluoroionophore, the absorption and emission of which undergo bathochromic shifts upon zinc(II) coordination. When the FRET donor is excited, efficient intramolecular <span class="hlt">energy</span> transfer occurs to result in the emission of the acceptor boron dipyrromethene (4,4-difluoro-4-bora-3a,4a-diaza-s-indacene or BODIPY) as a function of zinc(II) concentration. The broad emission <span class="hlt">band</span> of the donor/zinc(II) complex is transformed into the strong, narrow emission <span class="hlt">band</span> of the BODIPY acceptor in the FRET conjugates, which can be captured within the narrow emission window that is preferred for multicolor imaging experiments. In <span class="hlt">addition</span> to competing with other nonradiative decay processes of the FRET donor, the rapid intramolecular FRET of the excited FRET-conjugate molecule protects the donor fluorophore from photobleaching, thus enhancing the photostability of the indicator. FRET conjugates 3 and 4 contain aliphatic amino groups, which selectively target lysosomes in mammalian cells. This subcellular localization preference was verified by using confocal fluorescence microscopy, which also shows the zinc(II)-enhanced emission of 3 and 4 in lysosomes. It was further shown using two-color structured illumination microscopy (SIM), which is capable of extending the lateral resolution over the Abbe diffraction limit by a factor of two, that the morpholino-functionalized compound 4 localizes in the interior of lysosomes, rather than anchoring on the lysosomal membranes, of live HeLa cells. PMID:25382395</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015PhRvB..92l5441C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015PhRvB..92l5441C"><span><span class="hlt">Energy</span> shift and conduction-to-valence <span class="hlt">band</span> transition mediated by a time-dependent potential barrier in graphene</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Chaves, Andrey; da Costa, D. R.; de Sousa, G. O.; Pereira, J. M.; Farias, G. A.</p> <p>2015-09-01</p> <p>We investigate the scattering of a wave packet describing low-<span class="hlt">energy</span> electrons in graphene by a time-dependent finite-step potential barrier. Our results demonstrate that, after Klein tunneling through the barrier, the electron acquires an extra <span class="hlt">energy</span> which depends on the rate of change of the barrier height with time. If this rate is negative, the electron loses <span class="hlt">energy</span> and ends up as a valence <span class="hlt">band</span> state after leaving the barrier, which effectively behaves as a positively charged quasiparticle.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017PhRvB..95c5136D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017PhRvB..95c5136D"><span>Quadratic <span class="hlt">band</span> touching points and flat <span class="hlt">bands</span> in two-dimensional topological Floquet systems</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Du, Liang; Zhou, Xiaoting; Fiete, Gregory A.</p> <p>2017-01-01</p> <p>In this paper we theoretically study, using Floquet-Bloch theory, the influence of circularly and linearly polarized light on two-dimensional <span class="hlt">band</span> structures with Dirac and quadratic <span class="hlt">band</span> touching points, and flat <span class="hlt">bands</span>, taking the nearest neighbor hopping model on the kagome lattice as an example. We find circularly polarized light can invert the ordering of this three-<span class="hlt">band</span> model, while leaving the flat <span class="hlt">band</span> dispersionless. We find a small gap is also opened at the quadratic <span class="hlt">band</span> touching point by two-photon and higher order processes. By contrast, linearly polarized light splits the quadratic <span class="hlt">band</span> touching point (into two Dirac points) by an amount that depends only on the amplitude and polarization direction of the light, independent of the frequency, and generally renders dispersion to the flat <span class="hlt">band</span>. The splitting is perpendicular to the direction of the polarization of the light. We derive an effective low-<span class="hlt">energy</span> theory that captures these key results. Finally, we compute the frequency dependence of the optical conductivity for this three-<span class="hlt">band</span> model and analyze the various interband contributions of the Floquet modes. Our results suggest strategies for optically controlling <span class="hlt">band</span> structure and interaction strength in real systems.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li class="active"><span>7</span></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_7 --> <div id="page_8" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li class="active"><span>8</span></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="141"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/21985035','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21985035"><span>Toward tunable <span class="hlt">band</span> gap and tunable dirac point in bilayer graphene with molecular doping.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Yu, Woo Jong; Liao, Lei; Chae, Sang Hoon; Lee, Young Hee; Duan, Xiangfeng</p> <p>2011-11-09</p> <p>The bilayer graphene has attracted considerable attention for potential applications in future electronics and optoelectronics because of the feasibility to tune its <span class="hlt">band</span> gap with a vertical displacement field to break the inversion symmetry. Surface chemical doping in bilayer graphene can induce an <span class="hlt">additional</span> offset voltage to fundamentally affect the vertical displacement field and the <span class="hlt">band</span> gap opening in bilayer graphene. In this study, we investigate the effect of chemical molecular doping on <span class="hlt">band</span> gap opening in bilayer graphene devices with single or dual gate modulation. Chemical doping with benzyl viologen molecules modulates the displacement field to allow the opening of a transport <span class="hlt">band</span> gap and the increase of the on/off ratio in the bilayer graphene transistors. <span class="hlt">Additionally</span>, Fermi <span class="hlt">energy</span> level in the opened gap can be rationally controlled by the amount of molecular doping to obtain bilayer graphene transistors with tunable Dirac points, which can be readily configured into functional devices, such as complementary inverters.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1324969-new-silicon-phase-direct-band-gap-novel-optoelectronic-properties','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1324969-new-silicon-phase-direct-band-gap-novel-optoelectronic-properties"><span>A new silicon phase with direct <span class="hlt">band</span> gap and novel optoelectronic properties</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Guo, Yaguang; Wang, Qian; Kawazoe, Yoshiyuki; ...</p> <p>2015-09-23</p> <p>Due to the compatibility with the well-developed Si-based semiconductor industry, there is considerable interest in developing silicon structures with direct <span class="hlt">energy</span> <span class="hlt">band</span> gaps for effective sunlight harvesting. In this paper, using silicon triangles as the building block, we propose a new silicon allotrope with a direct <span class="hlt">band</span> gap of 0.61 eV, which is dynamically, thermally and mechanically stable. Symmetry group analysis further suggests that dipole transition at the direct <span class="hlt">band</span> gap is allowed. <span class="hlt">Additionally</span>, this new allotrope displays large carrier mobility (~10 4 cm/V · s) at room temperature and a low mass density (1.71 g/cm 3), making it amore » promising material for optoelectronic applications.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19800017026','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19800017026"><span>Shuttle Ku-<span class="hlt">band</span> and S-<span class="hlt">band</span> communications implementation study</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Dodds, J. G.; Huth, G. K.; Nilsen, P. W.; Polydoros, A.; Simon, M. K.; Weber, C. L.</p> <p>1980-01-01</p> <p>Various aspects of the shuttle orbiter S-<span class="hlt">band</span> network communication system, the S-<span class="hlt">band</span> payload communication system, and the Ku-<span class="hlt">band</span> communication system are considered. A method is proposed for obtaining more accurate S-<span class="hlt">band</span> antenna patterns of the actual shuttle orbiter vehicle during flight because the preliminary antenna patterns using mock-ups are not realistic that they do not include the effects of <span class="hlt">additional</span> appendages such as wings and tail structures. The Ku-<span class="hlt">band</span> communication system is discussed especially the TDRS antenna pointing accuracy with respect to the orbiter and the modifications required and resulting performance characteristics of the convolutionally encoded high data rate return link to maintain bit synchronizer lock on the ground. The TDRS user constraints on data bit clock jitter and data asymmetry on unbalanced QPSK with noisy phase references are included. The S-<span class="hlt">band</span> payload communication system study is outlined including the advantages and experimental results of a peak regulator design built and evaluated by Axiomatrix for the bent-pipe link versus the existing RMS-type regulator. The nominal sweep rate for the deep-space transponder of 250 Hz/s, and effects of phase noise on the performance of a communication system are analyzed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017PhRvL.119h7401Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017PhRvL.119h7401Y"><span>Optically Discriminating Carrier-Induced Quasiparticle <span class="hlt">Band</span> Gap and Exciton <span class="hlt">Energy</span> Renormalization in Monolayer MoS2</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yao, Kaiyuan; Yan, Aiming; Kahn, Salman; Suslu, Aslihan; Liang, Yufeng; Barnard, Edward S.; Tongay, Sefaattin; Zettl, Alex; Borys, Nicholas J.; Schuck, P. James</p> <p>2017-08-01</p> <p>Optoelectronic excitations in monolayer MoS2 manifest from a hierarchy of electrically tunable, Coulombic free-carrier and excitonic many-body phenomena. Investigating the fundamental interactions underpinning these phenomena—critical to both many-body physics exploration and device applications—presents challenges, however, due to a complex balance of competing optoelectronic effects and interdependent properties. Here, optical detection of bound- and free-carrier photoexcitations is used to directly quantify carrier-induced changes of the quasiparticle <span class="hlt">band</span> gap and exciton binding <span class="hlt">energies</span>. The results explicitly disentangle the competing effects and highlight longstanding theoretical predictions of large carrier-induced <span class="hlt">band</span> gap and exciton renormalization in two-dimensional semiconductors.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28952768','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28952768"><span>Optically Discriminating Carrier-Induced Quasiparticle <span class="hlt">Band</span> Gap and Exciton <span class="hlt">Energy</span> Renormalization in Monolayer MoS_{2}.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Yao, Kaiyuan; Yan, Aiming; Kahn, Salman; Suslu, Aslihan; Liang, Yufeng; Barnard, Edward S; Tongay, Sefaattin; Zettl, Alex; Borys, Nicholas J; Schuck, P James</p> <p>2017-08-25</p> <p>Optoelectronic excitations in monolayer MoS_{2} manifest from a hierarchy of electrically tunable, Coulombic free-carrier and excitonic many-body phenomena. Investigating the fundamental interactions underpinning these phenomena-critical to both many-body physics exploration and device applications-presents challenges, however, due to a complex balance of competing optoelectronic effects and interdependent properties. Here, optical detection of bound- and free-carrier photoexcitations is used to directly quantify carrier-induced changes of the quasiparticle <span class="hlt">band</span> gap and exciton binding <span class="hlt">energies</span>. The results explicitly disentangle the competing effects and highlight longstanding theoretical predictions of large carrier-induced <span class="hlt">band</span> gap and exciton renormalization in two-dimensional semiconductors.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018PhRvB..97t5113A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PhRvB..97t5113A"><span>Electronic and spin structure of the wide-<span class="hlt">band</span>-gap topological insulator: Nearly stoichiometric Bi2Te2S</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Annese, E.; Okuda, T.; Schwier, E. F.; Iwasawa, H.; Shimada, K.; Natamane, M.; Taniguchi, M.; Rusinov, I. P.; Eremeev, S. V.; Kokh, K. A.; Golyashov, V. A.; Tereshchenko, O. E.; Chulkov, E. V.; Kimura, A.</p> <p>2018-05-01</p> <p>We have grown the phase-homogeneous ternary compound with composition Bi2Te1.85S1.15 very close to the stoichiometric Bi2Te2S . The measurements performed with spin- and angle-resolved photoelectron spectroscopy as well as density functional theory and G W calculations revealed a wide-<span class="hlt">band</span>-gap three-dimensional topological insulator phase. The surface electronic spectrum is characterized by the topological surface state (TSS) with Dirac point located above the valence <span class="hlt">band</span> and Fermi level lying in the <span class="hlt">band</span> gap. TSS <span class="hlt">band</span> dispersion and constant <span class="hlt">energy</span> contour manifest a weak warping effect near the Fermi level along with in-plane and out-of-plane spin polarization along the Γ ¯-K ¯ line. We identified four <span class="hlt">additional</span> states at deeper binding <span class="hlt">energies</span> with high in-plane spin polarization.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1393189-shear-band-thickness-shear-band-cavities-zr-based-metallic-glass','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1393189-shear-band-thickness-shear-band-cavities-zr-based-metallic-glass"><span>Shear-<span class="hlt">band</span> thickness and shear-<span class="hlt">band</span> cavities in a Zr-based metallic glass</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Liu, C.; Roddatis, V.; Kenesei, P.; ...</p> <p>2017-08-14</p> <p>Strain localization into shear <span class="hlt">bands</span> in metallic glasses is typically described as a mechanism that occurs at the nano-scale, leaving behind a shear defect with a thickness of 10–20 nm. Here we sample the structure of a single system-spanning shear <span class="hlt">band</span> that has carried all plastic flow with high-angle annular dark field scanning transmission electron microscopy (HAADF-STEM) and high-<span class="hlt">energy</span> x-ray tomography (XRT). It is found that the shear-<span class="hlt">band</span> thickness and the density change relative to the matrix sensitively depend on position along the shear <span class="hlt">band</span>. A wide distribution of shear-<span class="hlt">band</span> thickness (10 nm–210 nm) and density change (–1% to –12%)more » is revealed. There is no obvious correlation between shear-<span class="hlt">band</span> thickness and density change, but larger thicknesses correspond typically to higher density changes. More than 100 micron-size shear-<span class="hlt">band</span> cavities were identified on the shear-<span class="hlt">band</span> plane, and their three-dimensional arrangement suggests a strongly fluctuating local curvature of the shear plane. As a result, these findings urge for a more complex view of a shear <span class="hlt">band</span> than a simple nano-scale planar defect.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/1393189-shear-band-thickness-shear-band-cavities-zr-based-metallic-glass','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/1393189-shear-band-thickness-shear-band-cavities-zr-based-metallic-glass"><span>Shear-<span class="hlt">band</span> thickness and shear-<span class="hlt">band</span> cavities in a Zr-based metallic glass</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Liu, C.; Roddatis, V.; Kenesei, P.</p> <p></p> <p>Strain localization into shear <span class="hlt">bands</span> in metallic glasses is typically described as a mechanism that occurs at the nano-scale, leaving behind a shear defect with a thickness of 10–20 nm. Here we sample the structure of a single system-spanning shear <span class="hlt">band</span> that has carried all plastic flow with high-angle annular dark field scanning transmission electron microscopy (HAADF-STEM) and high-<span class="hlt">energy</span> x-ray tomography (XRT). It is found that the shear-<span class="hlt">band</span> thickness and the density change relative to the matrix sensitively depend on position along the shear <span class="hlt">band</span>. A wide distribution of shear-<span class="hlt">band</span> thickness (10 nm–210 nm) and density change (–1% to –12%)more » is revealed. There is no obvious correlation between shear-<span class="hlt">band</span> thickness and density change, but larger thicknesses correspond typically to higher density changes. More than 100 micron-size shear-<span class="hlt">band</span> cavities were identified on the shear-<span class="hlt">band</span> plane, and their three-dimensional arrangement suggests a strongly fluctuating local curvature of the shear plane. As a result, these findings urge for a more complex view of a shear <span class="hlt">band</span> than a simple nano-scale planar defect.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25758749','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25758749"><span>Biologically inspired <span class="hlt">band</span>-edge laser action from semiconductor with dipole-forbidden <span class="hlt">band</span>-gap transition.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wang, Cih-Su; Liau, Chi-Shung; Sun, Tzu-Ming; Chen, Yu-Chia; Lin, Tai-Yuan; Chen, Yang-Fang</p> <p>2015-03-11</p> <p>A new approach is proposed to light up <span class="hlt">band</span>-edge stimulated emission arising from a semiconductor with dipole-forbidden <span class="hlt">band</span>-gap transition. To illustrate our working principle, here we demonstrate the feasibility on the composite of SnO2 nanowires (NWs) and chicken albumen. SnO2 NWs, which merely emit visible defect emission, are observed to generate a strong ultraviolet fluorescence centered at 387 nm assisted by chicken albumen at room temperature. In <span class="hlt">addition</span>, a stunning laser action is further discovered in the albumen/SnO2 NWs composite system. The underlying mechanism is interpreted in terms of the fluorescence resonance <span class="hlt">energy</span> transfer (FRET) from the chicken albumen protein to SnO2 NWs. More importantly, the giant oscillator strength of shallow defect states, which is served orders of magnitude larger than that of the free exciton, plays a decisive role. Our approach therefore shows that bio-materials exhibit a great potential in applications for novel light emitters, which may open up a new avenue for the development of bio-inspired optoelectronic devices.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2004APS..MARA15009C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2004APS..MARA15009C"><span>Role of Electronic Structure In Ion <span class="hlt">Band</span> State Theory of Low <span class="hlt">Energy</span> Nuclear Reactions</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Chubb, Scott</p> <p>2004-03-01</p> <p>The Nuts and Bolts of our Ion <span class="hlt">Band</span> State (IBS) theory of low <span class="hlt">energy</span> nuclear reactions (LENR's) in palladium-deuteride (PdD) and palladium-hydride (PdH) are the electrons that hold together or tear apart the bonds (or lack of bonds) between deuterons (d's) or protons (p's) and the host material. In PdDx and PdH_x, this bonding is strongly correlated with loading: in ambient loading conditions (x< 0. 6), the bonding in hibits IBS occupation. As x arrow 1, slight increases and decreases in loading can lead to vibrations (which have conventionally been thought to occur from phonons) that can induce potential losses or increases of p/d. Naive assumptions about phonons fail to include these losses and increases. These effects can occur because neither H or D has core electrons and because in either PdD or PdH, the electrons near the Fermi <span class="hlt">Energy</span> have negligible overlap with the nucleus of either D or H. I use these ideas to develop a formal justification, based on a generalization of conventional <span class="hlt">band</span> theory (Scott Chubb, "Semi-Classical Conduction of Charged and Neutral Particles in Finite Lattices," 2004 March Meeting."), for the idea that occupation of IBS's can occur and that this can lead to nuclear reactions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24104006','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24104006"><span>Electromagnetic waves in a topological insulator thin film stack: helicon-like wave mode and photonic <span class="hlt">band</span> structure.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Inoue, Jun-ichi</p> <p>2013-09-09</p> <p>We theoretically explore the electromagnetic modes specific to a topological insulator superlattice in which topological and conventional insulator thin films are stacked periodically. In particular, we obtain analytic formulas for low <span class="hlt">energy</span> mode that corresponds to a helicon wave, as well as those for photonic <span class="hlt">bands</span>. We illustrate that the system can be modeled as a stack of quantum Hall layers whose conductivity tensors alternately change signs, and then we analyze the photonic <span class="hlt">band</span> structures. This subject is a natural extension of a previous study by Tselis et al., which took into consideration a stack of identical quantum Hall layers but their discussion was limited into a low <span class="hlt">energy</span> mode. Thus we provide analytic formulas for photonic <span class="hlt">bands</span> and compare their features between the two systems. Our central findings in the topological insulator superlattice are that a low <span class="hlt">energy</span> mode corresponding to a helicon wave has linear dispersion instead of the conventional quadratic form, and that a robust gapless photonic <span class="hlt">band</span> appears although the system considered has spacial periodicity. In <span class="hlt">addition</span>, we demonstrate that the photonic <span class="hlt">bands</span> agree with the numerically calculated transmission spectra.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017APS..MARK13010D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017APS..MARK13010D"><span>Quadratic <span class="hlt">band</span> touching points and flat <span class="hlt">bands</span> in two-dimensional topological Floquet systems</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Du, Liang; Zhou, Xiaoting; Fiete, Gregory; The CenterComplex Quantum Systems Team</p> <p></p> <p>In this work we theoretically study, using Floquet-Bloch theory, the influence of circularly and linearly polarized light on two-dimensional <span class="hlt">band</span> structures with Dirac and quadratic <span class="hlt">band</span> touching points, and flat <span class="hlt">bands</span>, taking the nearest neighbor hopping model on the kagome lattice as an example. We find circularly polarized light can invert the ordering of this three <span class="hlt">band</span> model, while leaving the flat-<span class="hlt">band</span> dispersionless. We find a small gap is also opened at the quadratic <span class="hlt">band</span> touching point by 2-photon and higher order processes. By contrast, linearly polarized light splits the quadratic <span class="hlt">band</span> touching point (into two Dirac points) by an amount that depends only on the amplitude and polarization direction of the light, independent of the frequency, and generally renders dispersion to the flat <span class="hlt">band</span>. The splitting is perpendicular to the direction of the polarization of the light. We derive an effective low-<span class="hlt">energy</span> theory that captures these key results. Finally, we compute the frequency dependence of the optical conductivity for this 3-<span class="hlt">band</span> model and analyze the various interband contributions of the Floquet modes. Our results suggest strategies for optically controlling <span class="hlt">band</span> structure and interaction strength in real systems. We gratefully acknowledge funding from ARO Grant W911NF-14-1-0579 and NSF DMR-1507621.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017ApPRv...4b1301H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017ApPRv...4b1301H"><span><span class="hlt">Energy</span> <span class="hlt">band</span> offsets of dielectrics on InGaZnO4</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hays, David C.; Gila, B. P.; Pearton, S. J.; Ren, F.</p> <p>2017-06-01</p> <p>Thin-film transistors (TFTs) with channels made of hydrogenated amorphous silicon (a-Si:H) and polycrystalline silicon (poly-Si) are used extensively in the display industry. Amorphous silicon continues to dominate large-format display technology, but a-Si:H has a low electron mobility, μ ˜ 1 cm2/V s. Transparent, conducting metal-oxide materials such as Indium-Gallium-Zinc Oxide (IGZO) have demonstrated electron mobilities of 10-50 cm2/V s and are candidates to replace a-Si:H for TFT backplane technologies. The device performance depends strongly on the type of <span class="hlt">band</span> alignment of the gate dielectric with the semiconductor channel material and on the <span class="hlt">band</span> offsets. The factors that determine the conduction and valence <span class="hlt">band</span> offsets for a given material system are not well understood. Predictions based on various models have historically been unreliable and <span class="hlt">band</span> offset values must be determined experimentally. This paper provides experimental <span class="hlt">band</span> offset values for a number of gate dielectrics on IGZO for next generation TFTs. The relationship between <span class="hlt">band</span> offset and interface quality, as demonstrated experimentally and by previously reported results, is also explained. The literature shows significant variations in reported <span class="hlt">band</span> offsets and the reasons for these differences are evaluated. The biggest contributor to conduction <span class="hlt">band</span> offsets is the variation in the bandgap of the dielectrics due to differences in measurement protocols and stoichiometry resulting from different deposition methods, chemistry, and contamination. We have investigated the influence of valence <span class="hlt">band</span> offset values of strain, defects/vacancies, stoichiometry, chemical bonding, and contamination on IGZO/dielectric heterojunctions. These measurements provide data needed to further develop a predictive theory of <span class="hlt">band</span> offsets.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22483112-direct-band-gap-measurement-cu-ga-se-sub-thin-films-using-high-resolution-reflection-electron-energy-loss-spectroscopy','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22483112-direct-band-gap-measurement-cu-ga-se-sub-thin-films-using-high-resolution-reflection-electron-energy-loss-spectroscopy"><span>Direct <span class="hlt">band</span> gap measurement of Cu(In,Ga)(Se,S){sub 2} thin films using high-resolution reflection electron <span class="hlt">energy</span> loss spectroscopy</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Heo, Sung; College of Information and Communication Engineering, Sungkyunkwan University, Cheoncheon-dong 300, Jangan-gu, Suwon 440-746; Lee, Hyung-Ik</p> <p>2015-06-29</p> <p>To investigate the <span class="hlt">band</span> gap profile of Cu(In{sub 1−x},Ga{sub x})(Se{sub 1−y}S{sub y}){sub 2} of various compositions, we measured the <span class="hlt">band</span> gap profile directly as a function of in-depth using high-resolution reflection <span class="hlt">energy</span> loss spectroscopy (HR-REELS), which was compared with the <span class="hlt">band</span> gap profile calculated based on the auger depth profile. The <span class="hlt">band</span> gap profile is a double-graded <span class="hlt">band</span> gap as a function of in-depth. The calculated <span class="hlt">band</span> gap obtained from the auger depth profile seems to be larger than that by HR-REELS. Calculated <span class="hlt">band</span> gaps are to measure the average <span class="hlt">band</span> gap of the spatially different varying compositions with respectmore » to considering its void fraction. But, the results obtained using HR-REELS are to be affected by the low <span class="hlt">band</span> gap (i.e., out of void) rather than large one (i.e., near void). Our findings suggest an analytical method to directly determine the <span class="hlt">band</span> gap profile as function of in-depth.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018ChPhC..42e4104S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018ChPhC..42e4104S"><span><span class="hlt">Band</span> head spin assignment of superdeformed <span class="hlt">bands</span> in Hg isotopes through power index formula</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sharma, Honey; Mittal, H. M.</p> <p>2018-05-01</p> <p>The power index formula has been used to obtain the <span class="hlt">band</span> head spin (I 0) of all the superdeformed (SD) <span class="hlt">bands</span> in Hg isotopes. A least squares fitting approach is used. The root mean square deviations between the determined and the observed transition <span class="hlt">energies</span> are calculated by extracting the model parameters using the power index formula. Whenever definite spins are available, the determined and the observed transition <span class="hlt">energies</span> are in accordance with each other. The computed values of dynamic moment of inertia J (2) obtained by using the power index formula and its deviation with the rotational frequency is also studied. Excellent agreement is shown between the calculated and the experimental results for J (2) versus the rotational frequency. Hence, the power index formula works very well for all the SD <span class="hlt">bands</span> in Hg isotopes expect for 195Hg(2, 3, 4).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018PhRvL.120w7001L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PhRvL.120w7001L"><span>Electron Phonon Coupling versus Photoelectron <span class="hlt">Energy</span> Loss at the Origin of Replica <span class="hlt">Bands</span> in Photoemission of FeSe on SrTiO3</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Li, Fengmiao; Sawatzky, George A.</p> <p>2018-06-01</p> <p>The recent observation of replica <span class="hlt">bands</span> in single-layer FeSe /SrTiO3 by angle-resolved photoemission spectroscopy (ARPES) has triggered intense discussions concerning the potential influence of the FeSe electrons coupling with substrate phonons on the superconducting transition temperature. Here we provide strong evidence that the replica <span class="hlt">bands</span> observed in the single-layer FeSe /SrTiO3 system and several other cases are largely due to the <span class="hlt">energy</span> loss processes of the escaping photoelectron, resulted from the well-known strong coupling of external propagating electrons to Fuchs-Kliewer surface phonons in ionic materials in general. The photoelectron <span class="hlt">energy</span> loss in ARPES on single-layer FeSe /SrTiO3 is calculated using the demonstrated successful semiclassical dielectric theory in describing low <span class="hlt">energy</span> electron <span class="hlt">energy</span> loss spectroscopy of ionic insulators. Our result shows that the observed replica <span class="hlt">bands</span> are mostly a result of extrinsic photoelectron <span class="hlt">energy</span> loss and not a result of the electron phonon interaction of the Fe d electrons with the substrate phonons. The strong enhancement of the superconducting transition temperature in these monolayers remains an open question.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016PhRvB..93q4516C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016PhRvB..93q4516C"><span>Superconductivity versus bound-state formation in a two-<span class="hlt">band</span> superconductor with small Fermi <span class="hlt">energy</span>: Applications to Fe pnictides/chalcogenides and doped SrTiO3</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Chubukov, Andrey V.; Eremin, Ilya; Efremov, Dmitri V.</p> <p>2016-05-01</p> <p>We analyze the interplay between superconductivity and the formation of bound pairs of fermions (BCS-BEC crossover) in a 2D model of interacting fermions with small Fermi <span class="hlt">energy</span> EF and weak attractive interaction, which extends to <span class="hlt">energies</span> well above EF. The 2D case is special because a two-particle bound state forms at arbitrary weak interaction, and already at weak coupling, one has to distinguish between the bound-state formation and superconductivity. We briefly review the situation in the one-<span class="hlt">band</span> model and then consider two different two-<span class="hlt">band</span> models: one with one hole <span class="hlt">band</span> and one electron <span class="hlt">band</span> and another with two hole or two electron <span class="hlt">bands</span>. In each case, we obtain the bound-state <span class="hlt">energy</span> 2 E0 for two fermions in a vacuum and solve the set of coupled equations for the pairing gaps and the chemical potentials to obtain the onset temperature of the pairing Tins and the quasiparticle dispersion at T =0 . We then compute the superfluid stiffness ρs(T =0 ) and obtain the actual Tc. For definiteness, we set EF in one <span class="hlt">band</span> to be near zero and consider different ratios of E0 and EF in the other <span class="hlt">band</span>. We show that at EF≫E0 , the behavior of both two-<span class="hlt">band</span> models is BCS-like in the sense that Tc≈Tins≪EF and Δ ˜Tc . At EF≪E0 , the two models behave differently: in the model with two hole/two electron <span class="hlt">bands</span>, Tins˜E0/lnE/0EF , Δ ˜(E0EF) 1 /2 , and Tc˜EF , like in the one-<span class="hlt">band</span> model. In between Tins and Tc, the system displays a preformed pair behavior. In the model with one hole and one electron <span class="hlt">bands</span>, Tc remains of order Tins, and both remain finite at EF=0 and of the order of E0. The preformed pair behavior still does exist in this model because Tc is numerically smaller than Tins. For both models, we reexpress Tins in terms of the fully renormalized two-particle scattering amplitude by extending to the two-<span class="hlt">band</span> case (the method pioneered by Gorkov and Melik-Barkhudarov back in 1961). We apply our results for the model with a hole and an electron <span class="hlt">band</span> to</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018IJMPE..2750044G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018IJMPE..2750044G"><span>Novel solution of power law for γ-<span class="hlt">bands</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gupta, J. B.</p> <p></p> <p>The power law expression E = aIb offers a single-term formula with just two parameters for expressing the level <span class="hlt">energies</span> in the spectra of even-Z even-N nuclei. Its application to ground <span class="hlt">band</span> spectra for a wide range of nuclei has been demonstrated in our earlier works. Here, we extend its application to the rotational <span class="hlt">bands</span> built on an excited state of K = 2 γ-vibration <span class="hlt">band</span> and Kπ = 0 2+ beta <span class="hlt">band</span>. A novel assumption of a virtual level with spin zero for γ-<span class="hlt">bands</span> is made and its validity and use is illustrated. Here, the constancy of the parameters “b” and “a” with spin, offers a more realistic view of the dependence of the nuclear core deformation on spin, in the excited <span class="hlt">bands</span>. Also, it enables a spinwise view, not available in the other <span class="hlt">energy</span> fit expressions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018MolPh.116.1697J','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018MolPh.116.1697J"><span>Asymmetry induces Q-<span class="hlt">band</span> split in the electronic excitations of magnesium porphyrin</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Jiang, Xiankai; Gao, Yi; Lal, Ratnesh; Hu, Jun; Song, Bo</p> <p>2018-07-01</p> <p>The electronic excitations of magnesium porphyrin (MgP), a molecular model for understanding the physics in light harvesting by biological systems, have been studied extensively. However, the theoretical underpinning of experimental measurements is still lacking, especially about the sub-<span class="hlt">bands</span> in absorption spectrum. Here we propose that an asymmetry of MgP based on the uneven charge distribution of pyrrole rings and the linear structure of sp hybridised orbitals in Mg can largely influence the electronic excitations. Upon a very weak asymmetry of Mg-pyrrole bindings in MgP being introduced through the uneven distribution of charge, three different excitations are observed in the Q-<span class="hlt">band</span> region of the experimental spectrum. <span class="hlt">Additionally</span>, the predicted B-<span class="hlt">band</span> excitations are highly correlated (10-2 eV level) with experimental measurements. In contrast, without this asymmetry, there are only two degenerate excitations in the Q-<span class="hlt">band</span> region, and low agreement (10-1 eV level) of the B-<span class="hlt">band</span> excitations with the experiment. The key physics of the unexpected and observable asymmetry in MgP is the ability of Mg to form sp hybridised orbitals on the third shell upon Mg binding to the nitrogen of pyrrole ring. Our findings provide new insight for high-<span class="hlt">energy</span> efficiency of natural as well as artificial light-harvesting system for <span class="hlt">energy</span> challenge.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016SuMi...97..562O','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016SuMi...97..562O"><span>Twisted bilayer blue phosphorene: A direct <span class="hlt">band</span> gap semiconductor</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ospina, D. A.; Duque, C. A.; Correa, J. D.; Suárez Morell, Eric</p> <p>2016-09-01</p> <p>We report that two rotated layers of blue phosphorene behave as a direct <span class="hlt">band</span> gap semiconductor. The optical spectrum shows absorption peaks in the visible region of the spectrum and in <span class="hlt">addition</span> the <span class="hlt">energy</span> of these peaks can be tuned with the rotational angle. These findings makes twisted bilayer blue phosphorene a strong candidate as a solar cell or photodetection device. Our results are based on ab initio calculations of several rotated blue phosphorene layers.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li class="active"><span>8</span></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_8 --> <div id="page_9" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li class="active"><span>9</span></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="161"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/17995195','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/17995195"><span>NiO: correlated <span class="hlt">band</span> structure of a charge-transfer insulator.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kunes, J; Anisimov, V I; Skornyakov, S L; Lukoyanov, A V; Vollhardt, D</p> <p>2007-10-12</p> <p>The <span class="hlt">band</span> structure of the prototypical charge-transfer insulator NiO is computed by using a combination of an ab initio <span class="hlt">band</span> structure method and the dynamical mean-field theory with a quantum Monte-Carlo impurity solver. Employing a Hamiltonian which includes both Ni d and O p orbitals we find excellent agreement with the <span class="hlt">energy</span> <span class="hlt">bands</span> determined from angle-resolved photoemission spectroscopy. This brings an important progress in a long-standing problem of solid-state theory. Most notably we obtain the low-<span class="hlt">energy</span> Zhang-Rice <span class="hlt">bands</span> with strongly k-dependent orbital character discussed previously in the context of low-<span class="hlt">energy</span> model theories.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018SuMi..117..515Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018SuMi..117..515Z"><span>The localized effect of the Bi level on the valence <span class="hlt">band</span> in the dilute bismuth GaBixAs1-x alloy</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zhao, Chuan-Zhen; Zhu, Min-Min; Wang, Jun; Wang, Sha-Sha; Lu, Ke-Qing</p> <p>2018-05-01</p> <p>The research on the temperature dependence of the <span class="hlt">band</span> gap <span class="hlt">energy</span> of the dilute bismuth GaBixAs1-x alloy has been done. It is found that its temperature insensitiveness is due to the enhanced localized character of the valence <span class="hlt">band</span> state and the small decrease of the temperature coefficient for the conduction <span class="hlt">band</span> minimum (CBM). The enhanced localized character of the valence <span class="hlt">band</span> state is the main factor. In order to describe the localized effect of the Bi levels on the valence <span class="hlt">band</span>, the localized <span class="hlt">energy</span> is introduced into the Varshni's equation. It is found that the effect of the localized Bi level on the valence <span class="hlt">band</span> becomes strong with increasing Bi content. In <span class="hlt">addition</span>, it is found that the pressure dependence of the <span class="hlt">band</span> gap <span class="hlt">energy</span> of GaBixAs1-x does not seem to be influenced by the localized Bi levels. It is due to two factors. One is that the pressure dependence of the <span class="hlt">band</span> gap <span class="hlt">energy</span> is mainly determined by the D CBM of GaBixAs1-x. The D CBM of GaBixAs1-x is not influenced by the localized Bi levels. The other is that the small variation of the pressure coefficient for the D valence <span class="hlt">band</span> maximum (VBM) state of GaBixAs1-x can be cancelled by the variation of the pressure coefficient for the D CBM state of GaBixAs1-x.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22420794-study-energy-band-structure-photoelectrochemical-performances-spinel-li-sub-ti-sub-sub','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22420794-study-energy-band-structure-photoelectrochemical-performances-spinel-li-sub-ti-sub-sub"><span>Study on the <span class="hlt">energy</span> <span class="hlt">band</span> structure and photoelectrochemical performances of spinel Li{sub 4}Ti{sub 5}O{sub 12}</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Ge, Hao; Tian, Hui; Song, Hua</p> <p>2015-01-15</p> <p>Highlights: • Spinel Li{sub 4}Ti{sub 5}O{sub 12} possesses more positive potential of valence <span class="hlt">band</span> and wider <span class="hlt">band</span> gap than TiO{sub 2}. • Spinel Li{sub 4}Ti{sub 5}O{sub 12} displays typical n-type semiconductor characteristic and excellent UV-excitateded photocatalysis activity. • Our preliminary study will open new perspectives in investigation of other lithium-based compounds for new photocatalysts. - Abstract: <span class="hlt">Energy</span> <span class="hlt">band</span> structure, photoelectrochemical performances and photocatalysis activity of spinel Li{sub 4}Ti{sub 5}O{sub 12} are investigated for the first time in this paper. Li{sub 4}Ti{sub 5}O{sub 12} possesses more positive valence <span class="hlt">band</span> potential and wider <span class="hlt">band</span> gap than TiO{sub 2} due to its valencemore » <span class="hlt">band</span> consisting of Li{sub 1s} and Ti{sub 3d} orbitals mixed with O{sub 2p}. Li{sub 4}Ti{sub 5}O{sub 12} shows typical photocatalysis material characteristics and excellent photocatlytic activity under UV irradiation.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1990JCrGr.101..931B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1990JCrGr.101..931B"><span>Chemical trends of the luminescence in wide <span class="hlt">band</span> gap II 1-xMn xVI semimagnetic semiconductors</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Benecke, C.; Busse, W.; Gumlich, H.-E.</p> <p>1990-04-01</p> <p>Time resolved emission and excitation spectroscopy is used to investigate the Mn correlated luminescence in wide <span class="hlt">band</span> gap II-VI compounds, i.e. Zn 1-xMn xS, Cd 1-xMn xSe, Zn 1-xMn xTe and Cd 1-xMn xTe. <span class="hlt">Additional</span> Information has been obtained with CdxZnyMnzTe( x+ y+ z=1) in checking the luminescence by variation of the ratio of the cations Cd and Zn. Generally speaking, at least two distinct emissions <span class="hlt">bands</span> can be observed for each II 1- xMn xVI compound. One emissions <span class="hlt">band</span> is attributed to the internal transition 4T 1(G)→ 6A 1(S) of the 3d 5 electron of the Mn 2+ on regular metal sites with <span class="hlt">energies</span> of about ≈2 eV. The other emission <span class="hlt">band</span> is found to occur in the near infrared range of about ≈1.3 eV. This emission <span class="hlt">band</span> is tentatively interpreted as a transition of Mn 2+ ions on interstitial sites or in small Mn chalcogenide clusters, both interpretations assuming cubic symmetry. This model is supported by the existence of low <span class="hlt">energy</span> excitation <span class="hlt">bands</span> and by the great similarity of the shape of the two emission <span class="hlt">bands</span> which lead to comparable Huang-Rhys factors and effective phonon <span class="hlt">energies</span>. Also the established trend in the experimental data of the II-VI compounds under consideration confirm this interpretation. For both the IR and the yellow Mn 2+ center, the Racah parameters B and C and the crystal field parameter Dq are determined on the basis of experimental data. As a result, the <span class="hlt">energy</span> of both the emission and the excitation <span class="hlt">bands</span> is predominantly determined by the sorrounding anions. These <span class="hlt">bands</span> shift to higher <span class="hlt">energies</span> when the anions are changed in the fixed order: Te→Se→S. Regularly, there is also a spectral shift when Zn is replaced by Cd, which is smaller than the shift due to the variation of onions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28951053','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28951053"><span><span class="hlt">Additional</span> <span class="hlt">band</span> broadening of peptides in the first size-exclusion chromatographic dimension of an automated stop-flow two-dimensional high performance liquid chromatography.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Xu, Jucai; Sun-Waterhouse, Dongxiao; Qiu, Chaoying; Zhao, Mouming; Sun, Baoguo; Lin, Lianzhu; Su, Guowan</p> <p>2017-10-27</p> <p>The need to improve the peak capacity of liquid chromatography motivates the development of two-dimensional analysis systems. This paper presented a fully automated stop-flow two-dimensional liquid chromatography system with size exclusion chromatography followed by reversed phase liquid chromatography (SEC×RPLC) to efficiently separate peptides. The effects of different stop-flow operational parameters (stop-flow time, peak parking position, number of stop-flow periods and column temperature) on <span class="hlt">band</span> broadening in the first dimension (1 st D) SEC column were quantitatively evaluated by using commercial small proteins and peptides. Results showed that the effects of peak parking position and the number of stop-flow periods on <span class="hlt">band</span> broadening were relatively small. Unlike stop-flow analysis of large molecules with a long running time, <span class="hlt">additional</span> <span class="hlt">band</span> broadening was evidently observed for small molecule analytes due to the relatively high effective diffusion coefficient (D eff ). Therefore, shorter analysis time and lower 1 st D column temperature were suggested for analyzing small molecules. The stop-flow two-dimensional liquid chromatography (2D-LC) system was further tested on peanut peptides and an evidently improved resolution was observed for both stop-flow heart-cutting and comprehensive 2D-LC analysis (in spite of <span class="hlt">additional</span> <span class="hlt">band</span> broadening in SEC). The stop-flow SEC×RPLC, especially heart-cutting analysis with shorter analysis time and higher 1 st D resolution for selected fractions, offers a promising approach for efficient analysis of complex samples. Copyright © 2017 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/1429661-efficient-band-trap-tunneling-model-including-heterojunction-band-offset','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/1429661-efficient-band-trap-tunneling-model-including-heterojunction-band-offset"><span>Efficient <span class="hlt">Band</span>-to-Trap Tunneling Model Including Heterojunction <span class="hlt">Band</span> Offset</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Gao, Xujiao; Huang, Andy; Kerr, Bert</p> <p></p> <p>In this paper, we present an efficient <span class="hlt">band</span>-to-trap tunneling model based on the Schenk approach, in which an analytic density-of-states (DOS) model is developed based on the open boundary scattering method. The new model explicitly includes the effect of heterojunction <span class="hlt">band</span> offset, in <span class="hlt">addition</span> to the well-known field effect. Its analytic form enables straightforward implementation into TCAD device simulators. It is applicable to all one-dimensional potentials, which can be approximated to a good degree such that the approximated potentials lead to piecewise analytic wave functions with open boundary conditions. The model allows for simulating both the electric-field-enhanced and <span class="hlt">band</span>-offset-enhanced carriermore » recombination due to the <span class="hlt">band</span>-to-trap tunneling near the heterojunction in a heterojunction bipolar transistor (HBT). Simulation results of an InGaP/GaAs/GaAs NPN HBT show that the proposed model predicts significantly increased base currents, due to the hole-to-trap tunneling enhanced by the emitter-base junction <span class="hlt">band</span> offset. Finally, the results compare favorably with experimental observation.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1429661-efficient-band-trap-tunneling-model-including-heterojunction-band-offset','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1429661-efficient-band-trap-tunneling-model-including-heterojunction-band-offset"><span>Efficient <span class="hlt">Band</span>-to-Trap Tunneling Model Including Heterojunction <span class="hlt">Band</span> Offset</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Gao, Xujiao; Huang, Andy; Kerr, Bert</p> <p>2017-10-25</p> <p>In this paper, we present an efficient <span class="hlt">band</span>-to-trap tunneling model based on the Schenk approach, in which an analytic density-of-states (DOS) model is developed based on the open boundary scattering method. The new model explicitly includes the effect of heterojunction <span class="hlt">band</span> offset, in <span class="hlt">addition</span> to the well-known field effect. Its analytic form enables straightforward implementation into TCAD device simulators. It is applicable to all one-dimensional potentials, which can be approximated to a good degree such that the approximated potentials lead to piecewise analytic wave functions with open boundary conditions. The model allows for simulating both the electric-field-enhanced and <span class="hlt">band</span>-offset-enhanced carriermore » recombination due to the <span class="hlt">band</span>-to-trap tunneling near the heterojunction in a heterojunction bipolar transistor (HBT). Simulation results of an InGaP/GaAs/GaAs NPN HBT show that the proposed model predicts significantly increased base currents, due to the hole-to-trap tunneling enhanced by the emitter-base junction <span class="hlt">band</span> offset. Finally, the results compare favorably with experimental observation.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/10193446','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/10193446"><span>Table of superdeformed nuclear <span class="hlt">bands</span> and fission isomers</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Firestone, R.B.; Singh, B.</p> <p></p> <p>A minimum in the second potential well of deformed nuclei was predicted and the associated shell gaps are illustrated in the harmonic oscillator potential shell <span class="hlt">energy</span> surface calculations shown in this report. A strong superdeformed minimum in {sup 152}Dy was predicted for {beta}{sub 2}-0.65. Subsequently, a discrete set of {gamma}-ray transitions in {sup 152}DY was observed and, assigned to the predicted superdeformed <span class="hlt">band</span>. Extensive research at several laboratories has since focused on searching for other mass regions of large deformation. A new generation of {gamma}-ray detector arrays is already producing a wealth of information about the mechanisms for feeding andmore » deexciting superdeformed <span class="hlt">bands</span>. These <span class="hlt">bands</span> have been found in three distinct regions near A=l30, 150, and 190. This research extends upon previous work in the actinide region near A=240 where fission isomers were identified and also associated with the second potential well. Quadrupole moment measurements for selected cases in each mass region are consistent with assigning the <span class="hlt">bands</span> to excitations in the second local minimum. As part of our committment to maintain nuclear structure data as current as possible in the Evaluated Nuclear Structure Reference File (ENSDF) and the Table of Isotopes, we have updated the information on superdeformed nuclear <span class="hlt">bands</span>. As of April 1994, we have complied data from 86 superdeformed <span class="hlt">bands</span> and 46 fission isomers identified in 73 nuclides for this report. For each nuclide there is a complete level table listing both normal and superdeformed <span class="hlt">band</span> assignments; level <span class="hlt">energy</span>, spin, parity, half-life, magneto moments, decay branchings; and the <span class="hlt">energies</span>, final levels, relative intensities, multipolarities, and mixing ratios for transitions deexciting each level. Mass excess, decay <span class="hlt">energies</span>, and proton and neutron separation <span class="hlt">energies</span> are also provided from the evaluation of Audi and Wapstra.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JPSJ...87a4003M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JPSJ...87a4003M"><span>Size Reduction of Hamiltonian Matrix for Large-Scale <span class="hlt">Energy</span> <span class="hlt">Band</span> Calculations Using Plane Wave Bases</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Morifuji, Masato</p> <p>2018-01-01</p> <p>We present a method of reducing the size of a Hamiltonian matrix used in calculations of electronic states. In the electronic states calculations using plane wave basis functions, a large number of plane waves are often required to obtain precise results. Even using state-of-the-art techniques, the Hamiltonian matrix often becomes very large. The large computational time and memory necessary for diagonalization limit the widespread use of <span class="hlt">band</span> calculations. We show a procedure of deriving a reduced Hamiltonian constructed using a small number of low-<span class="hlt">energy</span> bases by renormalizing high-<span class="hlt">energy</span> bases. We demonstrate numerically that the significant speedup of eigenstates evaluation is achieved without losing accuracy.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015NIMPB.360..103C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015NIMPB.360..103C"><span><span class="hlt">Band</span> structure effects in the <span class="hlt">energy</span> loss of low-<span class="hlt">energy</span> protons and deuterons in thin films of Pt</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Celedón, C. E.; Sánchez, E. A.; Salazar Alarcón, L.; Guimpel, J.; Cortés, A.; Vargas, P.; Arista, N. R.</p> <p>2015-10-01</p> <p>We have investigated experimentally and by computer simulations the <span class="hlt">energy</span>-loss and angular distribution of low <span class="hlt">energy</span> (E < 10 keV) protons and deuterons transmitted through thin polycrystalline platinum films. The experimental results show significant deviations from the expected velocity dependence of the stopping power in the range of very low <span class="hlt">energies</span> with respect to the predictions of the Density Functional Theory for a jellium model. This behavior is similar to those observed in other transition metals such as Cu, Ag and Au, but different from the linear dependence recently observed in another transition metal, Pd, which belongs to the same Group of Pt in the Periodic Table. These differences are analyzed in term of the properties of the electronic <span class="hlt">bands</span> corresponding to Pt and Pd, represented in terms of the corresponding density of states. The present experiments include also a detailed study of the angular dependence of the <span class="hlt">energy</span> loss and the angular distributions of transmitted protons and deuterons. The results are compared with computer simulations based on the Monte Carlo method and with a theoretical model that evaluates the contributions of elastic collisions, path length effects in the inelastic <span class="hlt">energy</span> losses, and the effects of the foil roughness. The results of the analysis obtained from these various approaches provide a consistent and comprehensive description of the experimental findings.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011APS..MARP37005K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011APS..MARP37005K"><span>Spectromicroscopy measurements of surface morphology and <span class="hlt">band</span> structure of exfoliated graphene</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Knox, Kevin; Locatelli, Andrea; Cvetko, Dean; Mentes, Tevfik; Nino, Miguel; Wang, Shancai; Yilmaz, Mehmet; Kim, Philip; Osgood, Richard; Morgante, Alberto</p> <p>2011-03-01</p> <p>Monolayer-thick crystals, such as graphene, are an area of intense interest in condensed matter research. ~However, crystal deformations in these 2D systems are known to adversely affect conductivity and increase local chemical reactivity. <span class="hlt">Additionally</span>, surface roughness in graphene complicates <span class="hlt">band</span>-mapping and limits resolution in techniques such as angle resolved photoemission spectroscopy (ARPES), the theory of which was developed for atomically flat surfaces. Thus, an understanding of the surface morphology of graphene is essential to making high quality devices and important for interpreting ARPES results. In this talk, we will describe a non-invasive approach to examining the corrugation in exfoliated graphene using a combination of low <span class="hlt">energy</span> electron microscopy (LEEM) and micro-spot low <span class="hlt">energy</span> electron diffraction (LEED). We will also describe how such knowledge of surface roughness can be used in the analysis of ARPES data to improve resolution and extract useful information about the <span class="hlt">band</span>-structure.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018AAS...23145307A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AAS...23145307A"><span>High-<span class="hlt">energy</span> variability of the Pulsar binary PSR J1311-3430</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>An, Hongjun; Fermi-LAT Collaboration</p> <p>2018-01-01</p> <p>We present analysis results of high-<span class="hlt">energy</span> observations of the extreme mass-ratio black-widow millisecond pulsar binary PSR J1311-3430. Our studies in the UV, X-ray, and gamma-ray <span class="hlt">bands</span> confirm the orbital modulation in the gamma-ray <span class="hlt">band</span> as suggested previously. In <span class="hlt">addition</span>, we find that the modulation is stronger in the high-<span class="hlt">energy</span> <span class="hlt">band</span>. In the lower-<span class="hlt">energy</span> UV and X-ray <span class="hlt">bands</span>, we detect flares which were observed previously and attributed to magnetic activities. We find that the optical flares are associated with the X-ray flares, suggesting common origin. We explore possible connections of the variabilities with the intrabinary shock (IBS) and magnetic activity on the low mass companion.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26608712','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26608712"><span>Nanoscale measurements of unoccupied <span class="hlt">band</span> dispersion in few-layer graphene.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Jobst, Johannes; Kautz, Jaap; Geelen, Daniël; Tromp, Rudolf M; van der Molen, Sense Jan</p> <p>2015-11-26</p> <p>The properties of any material are fundamentally determined by its electronic <span class="hlt">band</span> structure. Each <span class="hlt">band</span> represents a series of allowed states inside a material, relating electron <span class="hlt">energy</span> and momentum. The occupied <span class="hlt">bands</span>, that is, the filled electron states below the Fermi level, can be routinely measured. However, it is remarkably difficult to characterize the empty part of the <span class="hlt">band</span> structure experimentally. Here, we present direct measurements of unoccupied <span class="hlt">bands</span> of monolayer, bilayer and trilayer graphene. To obtain these, we introduce a technique based on low-<span class="hlt">energy</span> electron microscopy. It relies on the dependence of the electron reflectivity on incidence angle and <span class="hlt">energy</span> and has a spatial resolution ∼10 nm. The method can be easily applied to other nanomaterials such as van der Waals structures that are available in small crystals only.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4674768','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4674768"><span>Nanoscale measurements of unoccupied <span class="hlt">band</span> dispersion in few-layer graphene</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Jobst, Johannes; Kautz, Jaap; Geelen, Daniël; Tromp, Rudolf M.; van der Molen, Sense Jan</p> <p>2015-01-01</p> <p>The properties of any material are fundamentally determined by its electronic <span class="hlt">band</span> structure. Each <span class="hlt">band</span> represents a series of allowed states inside a material, relating electron <span class="hlt">energy</span> and momentum. The occupied <span class="hlt">bands</span>, that is, the filled electron states below the Fermi level, can be routinely measured. However, it is remarkably difficult to characterize the empty part of the <span class="hlt">band</span> structure experimentally. Here, we present direct measurements of unoccupied <span class="hlt">bands</span> of monolayer, bilayer and trilayer graphene. To obtain these, we introduce a technique based on low-<span class="hlt">energy</span> electron microscopy. It relies on the dependence of the electron reflectivity on incidence angle and <span class="hlt">energy</span> and has a spatial resolution ∼10 nm. The method can be easily applied to other nanomaterials such as van der Waals structures that are available in small crystals only. PMID:26608712</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017SeScT..32i5002C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017SeScT..32i5002C"><span>Effect of conduction <span class="hlt">band</span> non-parabolicity on the optical gain of quantum cascade lasers based on the effective two-<span class="hlt">band</span> finite difference method</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Cho, Gookbin; Kim, Jungho</p> <p>2017-09-01</p> <p>We theoretically investigate the effect of conduction <span class="hlt">band</span> non-parabolicity (NPB) on the optical gain spectrum of quantum cascade lasers (QCLs) using the effective two-<span class="hlt">band</span> finite difference method. Based on the effective two-<span class="hlt">band</span> model to consider the NPB effect in the multiple quantum wells (QWs), the wave functions and confined <span class="hlt">energies</span> of electron states are calculated in two different active-region structures, which correspond to three-QW single-phonon and four-QW double-phonon resonance designs. In <span class="hlt">addition</span>, intersubband optical dipole moments and polar-optical-phonon scattering times are calculated and compared without and with the conduction <span class="hlt">band</span> NPB effect. Finally, the calculation results of optical gain spectra are compared in the two QCL structures having the same peak gain wavelength of 8.55 μm. The gain peaks are greatly shifted to longer wavelengths and the overall gain magnitudes are slightly reduced when the NPB effect is considered. Compared with the three-QW active-region design, the redshift of the peak gain is more prominent in the four-QW active-region design, which makes use of higher electronic states for the lasing transition.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25933339','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25933339"><span>Intermediate <span class="hlt">band</span> solar cell with extreme broadband spectrum quantum efficiency.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Datas, A; López, E; Ramiro, I; Antolín, E; Martí, A; Luque, A; Tamaki, R; Shoji, Y; Sogabe, T; Okada, Y</p> <p>2015-04-17</p> <p>We report, for the first time, about an intermediate <span class="hlt">band</span> solar cell implemented with InAs/AlGaAs quantum dots whose photoresponse expands from 250 to ∼6000  nm. To our knowledge, this is the broadest quantum efficiency reported to date for a solar cell and demonstrates that the intermediate <span class="hlt">band</span> solar cell is capable of producing photocurrent when illuminated with photons whose <span class="hlt">energy</span> equals the <span class="hlt">energy</span> of the lowest <span class="hlt">band</span> gap. We show experimental evidence indicating that this result is in agreement with the theory of the intermediate <span class="hlt">band</span> solar cell, according to which the generation recombination between the intermediate <span class="hlt">band</span> and the valence <span class="hlt">band</span> makes this photocurrent detectable.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017MsT.........13T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017MsT.........13T"><span><span class="hlt">Band</span> Gap Engineering of Titania Systems Purposed for Photocatalytic Activity</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Thurston, Cameron</p> <p></p> <p>Ab initio computer aided design drastically increases candidate population for highly specified material discovery and selection. These simulations, carried out through a first-principles computational approach, accurately extrapolate material properties and behavior. Titanium Dioxide (TiO2 ) is one such material that stands to gain a great deal from the use of these simulations. In its anatase form, titania (TiO2 ) has been found to exhibit a <span class="hlt">band</span> gap nearing 3.2 eV. If titania is to become a viable alternative to other contemporary photoactive materials exhibiting <span class="hlt">band</span> gaps better suited for the solar spectrum, then the <span class="hlt">band</span> gap must be subsequently reduced. To lower the <span class="hlt">energy</span> needed for electronic excitation, both transition metals and non-metals have been extensively researched and are currently viable candidates for the continued reduction of titania's <span class="hlt">band</span> gap. The introduction of multicomponent atomic doping introduces new <span class="hlt">energy</span> <span class="hlt">bands</span> which tend to both reduce the <span class="hlt">band</span> gap and recombination loss. Ta-N, Nb-N, V-N, Cr-N, Mo-N, and W-N substitutions were studied in titania and subsequent <span class="hlt">energy</span> and <span class="hlt">band</span> gap calculations show a favorable <span class="hlt">band</span> gap reduction in the case of passivated systems.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/15379532','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/15379532"><span>Relocation of the disulfonic stilbene sites of AE1 (<span class="hlt">band</span> 3) on the basis of fluorescence <span class="hlt">energy</span> transfer measurements.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Knauf, Philip A; Law, Foon-Yee; Leung, Tze-Wah Vivian; Atherton, Stephen J</p> <p>2004-09-28</p> <p>Previous fluorescence resonance <span class="hlt">energy</span> transfer (FRET) measurements, using BIDS (4-benzamido-4'-isothiocyanostilbene-2,2'-disulfonate) as a label for the disulfonic stilbene site and FM (fluorescein-5-maleimide) as a label for the cytoplasmic SH groups on <span class="hlt">band</span> 3 (AE1), combined with data showing that the cytoplasmic SH groups lie about 40 A from the cytoplasmic surface of the lipid bilayer, would place the BIDS sites very near the membrane's inner surface, a location that seems to be inconsistent with current models of AE1 structure and mechanism. We reinvestigated the BIDS-FM distance, using laser single photon counting techniques as well as steady-state fluorescence of AE1, in its native membrane environment. Both techniques agree that there is very little <span class="hlt">energy</span> transfer from BIDS to FM. The mean <span class="hlt">energy</span> transfer (E), based on three-exponential fits to the fluorescence decay data, is 2.5 +/- 0.7% (SEM, N = 12). Steady-state fluorescence measurements also indicate <3% <span class="hlt">energy</span> transfer from BIDS to FM. These data indicate that the BIDS sites are probably over 63 A from the cytoplasmic SH groups, placing them near the middle or the external half of the lipid bilayer. This relocation of the BIDS sites fits with other evidence that the disulfonic stilbene sites are located farther toward the external membrane surface than Glu-681, a residue near the inner membrane surface whose modification affects the pH dependence and anion selectivity of <span class="hlt">band</span> 3. The involvement of two relatively distant parts of the AE1 protein in transport function suggests that the transport mechanism requires coordinated large-scale conformational changes in the <span class="hlt">band</span> 3 protein.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018MPLB...3250355C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018MPLB...3250355C"><span>Atomistic full-quantum transport model for zigzag graphene nanoribbon-based structures: Complex <span class="hlt">energy-band</span> method</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Chen, Chun-Nan; Luo, Win-Jet; Shyu, Feng-Lin; Chung, Hsien-Ching; Lin, Chiun-Yan; Wu, Jhao-Ying</p> <p>2018-01-01</p> <p>Using a non-equilibrium Green’s function framework in combination with the complex <span class="hlt">energy-band</span> method, an atomistic full-quantum model for solving quantum transport problems for a zigzag-edge graphene nanoribbon (zGNR) structure is proposed. For transport calculations, the mathematical expressions from the theory for zGNR-based device structures are derived in detail. The transport properties of zGNR-based devices are calculated and studied in detail using the proposed method.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011BAAS...43..002C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011BAAS...43..002C"><span>Obituary: David L. <span class="hlt">Band</span> (1957-2009)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Cominsky, Lynn</p> <p>2011-12-01</p> <p> two new follow-up missions to CGRO, the Swift and Fermi observatories, <span class="hlt">Band</span> seized an opportunity in 2001 to join the staff of the Fermi Science Support Center at the NASA Goddard Space Flight Center in Greenbelt Maryland. He was hired as the lead scientist for user support functions and to help to define and implement planning for the 2008 launch of the Fermi spacecraft. He brought a high level of <span class="hlt">energy</span> and enthusiasm to the job, becoming in many ways the heart and soul of that organization. Neil Gehrels, the Goddard Astroparticle Physics Division Director and a Fermi deputy project scientist notes that "David was the perfect person for community support, with this outgoing personality and deep knowledge of astrophysics." <span class="hlt">Band</span> also became an important member of the Fermi science team; despite his failing health, he actively contributed to the first Fermi gamma-ray burst publication as well as making important contributions to the burst detection and data analysis techniques. <span class="hlt">Additionally</span>, <span class="hlt">Band</span> was known as a great communicator and mentor. He supervised a PhD student at UCSD who has subsequently been appointed to a faculty position. At Goddard, <span class="hlt">Band</span> was an integral part of the weekly scientific discussion groups within the gamma-ray astronomy group and he would always find the time to share his knowledge and expertise with new postdoctoral fellows and senior scientists alike. He was also involved with planning the EXIST mission, a candidate for a future NASA mission. He will be greatly missed by his many friends and colleagues within the Fermi mission and the high-<span class="hlt">energy</span> astrophysics community.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li class="active"><span>9</span></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_9 --> <div id="page_10" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li class="active"><span>10</span></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="181"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018AIPC.1920b0040S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AIPC.1920b0040S"><span>The effect of carbon nanotubes functionalization on the <span class="hlt">band</span>-gap <span class="hlt">energy</span> of TiO2-CNT nanocomposite</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Shahbazi, Hessam; Shafei, Alireza; Sheibani, Saeed</p> <p>2018-01-01</p> <p>In this paper the morphology and structure of TiO2-CNT nanocomposite powder obtained by an in situ sol-gel process were investigated. The synthesized nanocomposite powders were characterized by X-ray diffraction (XRD), field emission scanning electron microscopy (FESEM) and diffuse reflectance spectroscopy (DRS). The effect of functionalizing of CNT on the properties was studied. XRD results showed amorphous structure before calcination. Also, anatase phase TiO2 was formed after calcination at 400 °C. The SEM results indicate different distributions of TiO2 on CNTs. As a result, well dispersed TiO2 microstructure on the surface of CNTs was observed after functionalizing, while compact and large aggregated particles were found without functionalizing. The average thickness of uniform and well-defined coated TiO2 layer was in the range of 30-40 nm. The DRS results have determined the reflective properties and <span class="hlt">band</span> gap <span class="hlt">energies</span> of nanocomposite powders and have shown that functionalizing of CNTs caused the change of <span class="hlt">band</span>-gap <span class="hlt">energy</span> from 2.98 to 2.87 eV.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015OptEn..54i7102B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015OptEn..54i7102B"><span>Perfect metamaterial absorber-based <span class="hlt">energy</span> harvesting and sensor applications in the industrial, scientific, and medical <span class="hlt">band</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bakir, Mehmet; Karaaslan, Muharrem; Dincer, Furkan; Delihacioglu, Kemal; Sabah, Cumali</p> <p>2015-09-01</p> <p>An electromagnetic (EM) <span class="hlt">energy</span> harvesting application based on metamaterials is introduced. This application is operating at the the industrial, scientific, and medical <span class="hlt">band</span> (2.40 GHz), which is especially chosen because of its wide usage area. A square ring resonator (SRR) which has two gaps and two resistors across the gaps on it is used. Chip resistors are used to deliver the power to any active component that requires power. Transmission and reflection characteristics of the metamaterial absorber for <span class="hlt">energy</span> harvesting application are theoretically investigated and 83.6% efficient <span class="hlt">energy</span> harvesting application is realized. To prove that this study can be used for different sensor applications other than harvesting, a temperature sensor configuration is developed that can be applied to other sensing applications.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2010-title48-vol3/pdf/CFR-2010-title48-vol3-sec204-470.pdf','CFR'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2010-title48-vol3/pdf/CFR-2010-title48-vol3-sec204-470.pdf"><span>48 CFR 204.470 - U.S.-International Atomic <span class="hlt">Energy</span> Agency <span class="hlt">Additional</span> Protocol.</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collectionCfr.action?selectedYearFrom=2010&page.go=Go">Code of Federal Regulations, 2010 CFR</a></p> <p></p> <p>2010-10-01</p> <p>... 48 Federal Acquisition Regulations System 3 2010-10-01 2010-10-01 false U.S.-International Atomic <span class="hlt">Energy</span> Agency <span class="hlt">Additional</span> Protocol. 204.470 Section 204.470 Federal Acquisition Regulations System DEFENSE... Information Within Industry 204.470 U.S.-International Atomic <span class="hlt">Energy</span> Agency <span class="hlt">Additional</span> Protocol. ...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2011-title48-vol3/pdf/CFR-2011-title48-vol3-sec204-470.pdf','CFR2011'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2011-title48-vol3/pdf/CFR-2011-title48-vol3-sec204-470.pdf"><span>48 CFR 204.470 - U.S.-International Atomic <span class="hlt">Energy</span> Agency <span class="hlt">Additional</span> Protocol.</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collectionCfr.action?selectedYearFrom=2011&page.go=Go">Code of Federal Regulations, 2011 CFR</a></p> <p></p> <p>2011-10-01</p> <p>... 48 Federal Acquisition Regulations System 3 2011-10-01 2011-10-01 false U.S.-International Atomic <span class="hlt">Energy</span> Agency <span class="hlt">Additional</span> Protocol. 204.470 Section 204.470 Federal Acquisition Regulations System DEFENSE... Information Within Industry 204.470 U.S.-International Atomic <span class="hlt">Energy</span> Agency <span class="hlt">Additional</span> Protocol. ...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2012-title48-vol3/pdf/CFR-2012-title48-vol3-sec204-470.pdf','CFR2012'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2012-title48-vol3/pdf/CFR-2012-title48-vol3-sec204-470.pdf"><span>48 CFR 204.470 - U.S.-International Atomic <span class="hlt">Energy</span> Agency <span class="hlt">Additional</span> Protocol.</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collectionCfr.action?selectedYearFrom=2012&page.go=Go">Code of Federal Regulations, 2012 CFR</a></p> <p></p> <p>2012-10-01</p> <p>... 48 Federal Acquisition Regulations System 3 2012-10-01 2012-10-01 false U.S.-International Atomic <span class="hlt">Energy</span> Agency <span class="hlt">Additional</span> Protocol. 204.470 Section 204.470 Federal Acquisition Regulations System DEFENSE... Information Within Industry 204.470 U.S.-International Atomic <span class="hlt">Energy</span> Agency <span class="hlt">Additional</span> Protocol. ...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2014-title48-vol3/pdf/CFR-2014-title48-vol3-sec204-470.pdf','CFR2014'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2014-title48-vol3/pdf/CFR-2014-title48-vol3-sec204-470.pdf"><span>48 CFR 204.470 - U.S.-International Atomic <span class="hlt">Energy</span> Agency <span class="hlt">Additional</span> Protocol.</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collectionCfr.action?selectedYearFrom=2014&page.go=Go">Code of Federal Regulations, 2014 CFR</a></p> <p></p> <p>2014-10-01</p> <p>... 48 Federal Acquisition Regulations System 3 2014-10-01 2014-10-01 false U.S.-International Atomic <span class="hlt">Energy</span> Agency <span class="hlt">Additional</span> Protocol. 204.470 Section 204.470 Federal Acquisition Regulations System DEFENSE... Information Within Industry 204.470 U.S.-International Atomic <span class="hlt">Energy</span> Agency <span class="hlt">Additional</span> Protocol. ...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2013-title48-vol3/pdf/CFR-2013-title48-vol3-sec204-470.pdf','CFR2013'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2013-title48-vol3/pdf/CFR-2013-title48-vol3-sec204-470.pdf"><span>48 CFR 204.470 - U.S.-International Atomic <span class="hlt">Energy</span> Agency <span class="hlt">Additional</span> Protocol.</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collectionCfr.action?selectedYearFrom=2013&page.go=Go">Code of Federal Regulations, 2013 CFR</a></p> <p></p> <p>2013-10-01</p> <p>... 48 Federal Acquisition Regulations System 3 2013-10-01 2013-10-01 false U.S.-International Atomic <span class="hlt">Energy</span> Agency <span class="hlt">Additional</span> Protocol. 204.470 Section 204.470 Federal Acquisition Regulations System DEFENSE... Information Within Industry 204.470 U.S.-International Atomic <span class="hlt">Energy</span> Agency <span class="hlt">Additional</span> Protocol. ...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016APS..MARH47005T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016APS..MARH47005T"><span>New insights into the opening <span class="hlt">band</span> gap of graphene oxides</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Tran, Ngoc Thanh Thuy; Lin, Shih-Yang; Lin, Ming-Fa</p> <p></p> <p>Electronic properties of oxygen absorbed few-layer graphenes are investigated using first-principle calculations. They are very sensitive to the changes in the oxygen concentration, number of graphene layer, and stacking configuration. The feature-rich <span class="hlt">band</span> structures exhibit the destruction or distortion of the Dirac cone, opening of <span class="hlt">band</span> gap, anisotropic <span class="hlt">energy</span> dispersions, O- and (C,O)-dominated <span class="hlt">energy</span> dispersions, and extra critical points. The <span class="hlt">band</span> decomposed charge distributions reveal the π-bonding dominated <span class="hlt">energy</span> gap. The orbital-projected density of states (DOS) have many special structures mainly coming from a composite <span class="hlt">energy</span> <span class="hlt">band</span>, the parabolic and partially flat ones. The DOS and spatial charge distributions clearly indicate the critical orbital hybridizations in O-O, C-O and C-C bonds, being responsible for the diversified properties. All of the few-layer graphene oxides are semi-metals except for the semiconducting monolayer ones.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22606440-energy-band-gap-spectroscopic-studies-mn-sub-cu-sub-wo-sub','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22606440-energy-band-gap-spectroscopic-studies-mn-sub-cu-sub-wo-sub"><span><span class="hlt">Energy</span> <span class="hlt">band</span> gap and spectroscopic studies in Mn{sub 1-x}Cu{sub x}WO{sub 4} (0 ≤ x ≤ 0.125)</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Mal, Priyanath; Rambabu, P.; Turpu, G. R.</p> <p>2016-05-06</p> <p>A study on the effect of nonmagnetic Cu{sup 2+} substitution at Mn{sup 2+} site on the structural and <span class="hlt">energy</span> <span class="hlt">band</span> gap of the MnWO{sub 4} is reported. Convenient solid state reaction route has been adopted for the synthesis of Mn{sub 1-x}Cu{sub x}WO{sub 4}. X-ray diffraction (XRD) pattern showed high crystalline quality of the prepared samples. Raman spectroscopic studies were carried out to understand the structural aspects of the doping. 15 Raman active modes were identified out of 18, predicted for wolframite type monoclinic structure of MnWO{sub 4}. UV-visible diffuse reflectance spectra were recorded and analyzed to get <span class="hlt">energy</span> <span class="hlt">band</span> gapmore » of the studied system and are found in the range of 2.5 eV to 2.04 eV with a systematic decrease with the increase in Cu{sup 2+} concentration. <span class="hlt">Energy</span> <span class="hlt">band</span> gap values are verified by Density Functional Theory calculations based on projector augmented wave (PAW) method. The calculated values are in good agreement with the experimental data.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22493114-measurement-inassb-bandgap-energy-inas-inassb-band-edge-positions-using-spectroscopic-ellipsometry-photoluminescence-spectroscopy','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22493114-measurement-inassb-bandgap-energy-inas-inassb-band-edge-positions-using-spectroscopic-ellipsometry-photoluminescence-spectroscopy"><span>Measurement of InAsSb bandgap <span class="hlt">energy</span> and InAs/InAsSb <span class="hlt">band</span> edge positions using spectroscopic ellipsometry and photoluminescence spectroscopy</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Webster, P. T.; Riordan, N. A.; Liu, S.</p> <p>2015-12-28</p> <p>The structural and optical properties of lattice-matched InAs{sub 0.911}Sb{sub 0.089} bulk layers and strain-balanced InAs/InAs{sub 1−x}Sb{sub x} (x ∼ 0.1–0.4) superlattices grown on (100)-oriented GaSb substrates by molecular beam epitaxy are examined using X-ray diffraction, spectroscopic ellipsometry, and temperature dependent photoluminescence spectroscopy. The photoluminescence and ellipsometry measurements determine the ground state bandgap <span class="hlt">energy</span> and the X-ray diffraction measurements determine the layer thickness and mole fraction of the structures studied. Detailed modeling of the X-ray diffraction data is employed to quantify unintentional incorporation of approximately 1% Sb into the InAs layers of the superlattices. A Kronig-Penney model of the superlattice miniband structure ismore » used to analyze the valence <span class="hlt">band</span> offset between InAs and InAsSb, and hence the InAsSb <span class="hlt">band</span> edge positions at each mole fraction. The resulting composition dependence of the bandgap <span class="hlt">energy</span> and <span class="hlt">band</span> edge positions of InAsSb are described using the bandgap bowing model; the respective low and room temperature bowing parameters for bulk InAsSb are 938 and 750 meV for the bandgap, 558 and 383 meV for the conduction <span class="hlt">band</span>, and −380 and −367 meV for the valence <span class="hlt">band</span>.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016JMoSp.324...12Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016JMoSp.324...12Z"><span>Direct measurement of <span class="hlt">additional</span> Ar-H2O vibration-rotation-tunneling <span class="hlt">bands</span> in the millimeter-submillimeter range</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zou, Luyao; Widicus Weaver, Susanna L.</p> <p>2016-06-01</p> <p>Three new weak <span class="hlt">bands</span> of the Ar-H2O vibration-rotation-tunneling spectrum have been measured in the millimeter wavelength range. These <span class="hlt">bands</span> were predicted from combination differences based on previously measured <span class="hlt">bands</span> in the submillimeter region. Two previously reported submillimeter <span class="hlt">bands</span> were also remeasured with higher frequency resolution. These new measurements allow us to obtain accurate information on the Coriolis interaction between the 101 and 110 states. Here we report these results and the associated improved molecular constants.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29484486','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29484486"><span>Computational Design of Flat-<span class="hlt">Band</span> Material.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Hase, I; Yanagisawa, T; Kawashima, K</p> <p>2018-02-26</p> <p>Quantum mechanics states that hopping integral between local orbitals makes the <span class="hlt">energy</span> <span class="hlt">band</span> dispersive. However, in some special cases, there are <span class="hlt">bands</span> with no dispersion due to quantum interference. These <span class="hlt">bands</span> are called as flat <span class="hlt">band</span>. Many models having flat <span class="hlt">band</span> have been proposed, and many interesting physical properties are predicted. However, no real compound having flat <span class="hlt">band</span> has been found yet despite the 25 years of vigorous researches. We have found that some pyrochlore oxides have quasi-flat <span class="hlt">band</span> just below the Fermi level by first principles calculation. Moreover, their valence <span class="hlt">bands</span> are well described by a tight-binding model of pyrochlore lattice with isotropic nearest neighbor hopping integral. This model belongs to a class of Mielke model, whose ground state is known to be ferromagnetic with appropriate carrier doping and on-site repulsive Coulomb interaction. We have also performed a spin-polarized <span class="hlt">band</span> calculation for the hole-doped system from first principles and found that the ground state is ferromagnetic for some doping region. Interestingly, these compounds do not include magnetic element, such as transition metal and rare-earth elements.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018NRL....13...63H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018NRL....13...63H"><span>Computational Design of Flat-<span class="hlt">Band</span> Material</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hase, I.; Yanagisawa, T.; Kawashima, K.</p> <p>2018-02-01</p> <p>Quantum mechanics states that hopping integral between local orbitals makes the <span class="hlt">energy</span> <span class="hlt">band</span> dispersive. However, in some special cases, there are <span class="hlt">bands</span> with no dispersion due to quantum interference. These <span class="hlt">bands</span> are called as flat <span class="hlt">band</span>. Many models having flat <span class="hlt">band</span> have been proposed, and many interesting physical properties are predicted. However, no real compound having flat <span class="hlt">band</span> has been found yet despite the 25 years of vigorous researches. We have found that some pyrochlore oxides have quasi-flat <span class="hlt">band</span> just below the Fermi level by first principles calculation. Moreover, their valence <span class="hlt">bands</span> are well described by a tight-binding model of pyrochlore lattice with isotropic nearest neighbor hopping integral. This model belongs to a class of Mielke model, whose ground state is known to be ferromagnetic with appropriate carrier doping and on-site repulsive Coulomb interaction. We have also performed a spin-polarized <span class="hlt">band</span> calculation for the hole-doped system from first principles and found that the ground state is ferromagnetic for some doping region. Interestingly, these compounds do not include magnetic element, such as transition metal and rare-earth elements.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19750050946&hterms=nike&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dnike','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19750050946&hterms=nike&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dnike"><span>Electron currents associated with an auroral <span class="hlt">band</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Spiger, R. J.; Anderson, H. R.</p> <p>1975-01-01</p> <p>Measurements of electron pitch angle distributions and <span class="hlt">energy</span> spectra over a broad auroral <span class="hlt">band</span> were used to calculate net electric current carried by auroral electrons in the vicinity of the <span class="hlt">band</span>. The particle <span class="hlt">energy</span> spectrometers were carried by a Nike-Tomahawk rocket launched from Poker Flat, Alaska, at 0722 UT on February 25, 1972. Data are presented which indicate the existence of upward field-aligned currents of electrons in the <span class="hlt">energy</span> range 0.5-20 keV. The spatial relationship of these currents to visual structure of the auroral arc and the characteristics of the electrons carrying the currents are discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1259474-evidence-shockley-read-hall-defect-state-independent-band-edge-energy-inas-sb-type-ii-superlattices','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1259474-evidence-shockley-read-hall-defect-state-independent-band-edge-energy-inas-sb-type-ii-superlattices"><span>Evidence of a Shockley-Read-Hall Defect State Independent of <span class="hlt">Band</span>-Edge <span class="hlt">Energy</span> in InAs / In ( As , Sb ) Type-II Superlattices</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Aytac, Y.; Olson, B. V.; Kim, J. K.; ...</p> <p>2016-06-01</p> <p>A set of seven InAs/InAsSb type-II superlattices (T2SLs) were designed to have speci c bandgap <span class="hlt">energies</span> between 290 meV (4.3 m) and 135 meV (9.2 m) in order to study the e ects of the T2SL bandgap <span class="hlt">energy</span> on the minority carrier lifetime. A temperature dependent optical pump-probe technique is used to measure the carrier lifetimes, and the e ect of a mid-gap defect level on the carrier recombination dynamics is reported. The Shockley-Read-Hall (SRH) defect state is found to be at <span class="hlt">energy</span> of approximately -250 12 meV relative to the valence <span class="hlt">band</span> edge of bulk GaSb for the entiremore » set of T2SL structures, even though the T2SL valence <span class="hlt">band</span> edge shifts by 155 meV on the same scale. These results indicate that the SRH defect state in InAs/InAsSb T2SLs is singular and is nearly independent of the exact position of the T2SL bandgap or <span class="hlt">band</span> edge <span class="hlt">energies</span>. They also suggest the possibility of engineering the T2SL structure such that the SRH state is removed completely from the bandgap, a result that should signi cantly increase the minority carrier lifetime.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1259474','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1259474"><span>Evidence of a Shockley-Read-Hall Defect State Independent of <span class="hlt">Band</span>-Edge <span class="hlt">Energy</span> in InAs / In ( As , Sb ) Type-II Superlattices</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Aytac, Y.; Olson, B. V.; Kim, J. K.</p> <p></p> <p>A set of seven InAs/InAsSb type-II superlattices (T2SLs) were designed to have speci c bandgap <span class="hlt">energies</span> between 290 meV (4.3 m) and 135 meV (9.2 m) in order to study the e ects of the T2SL bandgap <span class="hlt">energy</span> on the minority carrier lifetime. A temperature dependent optical pump-probe technique is used to measure the carrier lifetimes, and the e ect of a mid-gap defect level on the carrier recombination dynamics is reported. The Shockley-Read-Hall (SRH) defect state is found to be at <span class="hlt">energy</span> of approximately -250 12 meV relative to the valence <span class="hlt">band</span> edge of bulk GaSb for the entiremore » set of T2SL structures, even though the T2SL valence <span class="hlt">band</span> edge shifts by 155 meV on the same scale. These results indicate that the SRH defect state in InAs/InAsSb T2SLs is singular and is nearly independent of the exact position of the T2SL bandgap or <span class="hlt">band</span> edge <span class="hlt">energies</span>. They also suggest the possibility of engineering the T2SL structure such that the SRH state is removed completely from the bandgap, a result that should signi cantly increase the minority carrier lifetime.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1022533','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1022533"><span>Insights on the Cuprate High <span class="hlt">Energy</span> Anomaly Observed in ARPES</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Moritz, Brian</p> <p>2011-08-16</p> <p>Recently, angle-resolved photoemission spectroscopy has been used to highlight an anomalously large <span class="hlt">band</span> renormalization at high binding <span class="hlt">energies</span> in cuprate superconductors: the high <span class="hlt">energy</span> 'waterfall' or high <span class="hlt">energy</span> anomaly (HEA). The anomaly is present for both hole- and electron-doped cuprates as well as the half-filled parent insulators with different <span class="hlt">energy</span> scales arising on either side of the phase diagram. While photoemission matrix elements clearly play a role in changing the aesthetic appearance of the <span class="hlt">band</span> dispersion, i.e. creating a 'waterfall'-like appearance, they provide an inadequate description for the physics that underlies the strong <span class="hlt">band</span> renormalization giving rise to the HEA.more » Model calculations of the single-<span class="hlt">band</span> Hubbard Hamiltonian showcase the role played by correlations in the formation of the HEA and uncover significant differences in the HEA <span class="hlt">energy</span> scale for hole- and electron-doped cuprates. In <span class="hlt">addition</span>, this approach properly captures the transfer of spectral weight accompanying doping in a correlated material and provides a unifying description of the HEA across both sides of the cuprate phase diagram. We find that the anomaly demarcates a transition, or cross-over, from a quasiparticle <span class="hlt">band</span> at low binding <span class="hlt">energies</span> near the Fermi level to valence <span class="hlt">bands</span> at higher binding <span class="hlt">energy</span>, assumed to be of strong oxygen character.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.dtic.mil/docs/citations/ADA625566','DTIC-ST'); return false;" href="http://www.dtic.mil/docs/citations/ADA625566"><span>Conduction- and Valence-<span class="hlt">Band</span> <span class="hlt">Energies</span> in Bulk InAs(1-x)Sb(x) and Type II InAs(1-x) Sb(x)/InAs Strained-Layer Superlattices</span></a></p> <p><a target="_blank" href="http://www.dtic.mil/">DTIC Science & Technology</a></p> <p></p> <p>2013-03-08</p> <p>tions in the studied SLS structures . The fit of the dependence of the valence- <span class="hlt">band</span> <span class="hlt">energy</span> of unstrained InAs1!xSbx on the composition x with a... <span class="hlt">band</span> . STRUCTURES Bulk InAsSb epilayers on metamorphic buffers and InAsSb/InAs strained-layer superlattices (SLS) were grown on GaSb substrates by solid...meV in InAs and Ev = 0 meV in InSb. For InAsSb with 22.5% Sb grown on GaSb , an unstrained valence- <span class="hlt">band</span> <span class="hlt">energy</span> of Ev = !457 meV was obtained. For the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016OptCo.367..192Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016OptCo.367..192Z"><span>Accurate <span class="hlt">band-to-band</span> registration of AOTF imaging spectrometer using motion detection technology</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zhou, Pengwei; Zhao, Huijie; Jin, Shangzhong; Li, Ningchuan</p> <p>2016-05-01</p> <p>This paper concerns the problem of platform vibration induced <span class="hlt">band-to-band</span> misregistration with acousto-optic imaging spectrometer in spaceborne application. Registrating images of different <span class="hlt">bands</span> formed at different time or different position is difficult, especially for hyperspectral images form acousto-optic tunable filter (AOTF) imaging spectrometer. In this study, a motion detection method is presented using the polychromatic undiffracted beam of AOTF. The factors affecting motion detect accuracy are analyzed theoretically, and calculations show that optical distortion is an easily overlooked factor to achieve accurate <span class="hlt">band-to-band</span> registration. Hence, a reflective dual-path optical system has been proposed for the first time, with reduction of distortion and chromatic aberration, indicating the potential of higher registration accuracy. Consequently, a spectra restoration experiment using <span class="hlt">additional</span> motion detect channel is presented for the first time, which shows the accurate spectral image registration capability of this technique.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012EPJB...85..324G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012EPJB...85..324G"><span>Plasmon satellites in valence-<span class="hlt">band</span> photoemission spectroscopy. Ab initio study of the photon-<span class="hlt">energy</span> dependence in semiconductors</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Guzzo, M.; Kas, J. J.; Sottile, F.; Silly, M. G.; Sirotti, F.; Rehr, J. J.; Reining, L.</p> <p>2012-09-01</p> <p>We present experimental data and theoretical results for valence-<span class="hlt">band</span> satellites in semiconductors, using the prototypical example of bulk silicon. In a previous publication we introduced a new approach that allows us to describe satellites in valence photoemission spectroscopy, in good agreement with experiment. Here we give more details; we show how the the spectra change with photon <span class="hlt">energy</span>, and how the theory explains this behaviour. We also describe how we include several effects which are important to obtain a correct comparison between theory and experiment, such as secondary electrons and photon cross sections. In particular the inclusion of extrinsic losses and their dependence on the photon <span class="hlt">energy</span> are key to the description of the <span class="hlt">energy</span> dependence of spectra.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li class="active"><span>10</span></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_10 --> <div id="page_11" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li class="active"><span>11</span></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="201"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018PhRvB..97c5138B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PhRvB..97c5138B"><span><span class="hlt">Band</span> connectivity for topological quantum chemistry: <span class="hlt">Band</span> structures as a graph theory problem</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bradlyn, Barry; Elcoro, L.; Vergniory, M. G.; Cano, Jennifer; Wang, Zhijun; Felser, C.; Aroyo, M. I.; Bernevig, B. Andrei</p> <p>2018-01-01</p> <p>The conventional theory of solids is well suited to describing <span class="hlt">band</span> structures locally near isolated points in momentum space, but struggles to capture the full, global picture necessary for understanding topological phenomena. In part of a recent paper [B. Bradlyn et al., Nature (London) 547, 298 (2017), 10.1038/nature23268], we have introduced the way to overcome this difficulty by formulating the problem of sewing together many disconnected local k .p <span class="hlt">band</span> structures across the Brillouin zone in terms of graph theory. In this paper, we give the details of our full theoretical construction. We show that crystal symmetries strongly constrain the allowed connectivities of <span class="hlt">energy</span> <span class="hlt">bands</span>, and we employ graph theoretic techniques such as graph connectivity to enumerate all the solutions to these constraints. The tools of graph theory allow us to identify disconnected groups of <span class="hlt">bands</span> in these solutions, and so identify topologically distinct insulating phases.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23484179','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23484179"><span>Do adhesive systems leave resin coats on the surfaces of the metal matrix <span class="hlt">bands</span>? An adhesive remnant characterization.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Arhun, Neslihan; Cehreli, Sevi Burcak</p> <p>2013-01-01</p> <p>Reestablishing proximal contacts with composite resins may prove challenging since the applied adhesives may lead to resin coating that produces <span class="hlt">additional</span> thickness. The aim of this study was to investigate the surface of metal matrix <span class="hlt">bands</span> after application of adhesive systems and blowing or wiping off the adhesive before polymerization. Seventeen groups of matrix <span class="hlt">bands</span> were prepared. The remnant particles were characterized by <span class="hlt">energy</span> dispersive spectrum and scanning electron microscopy. Total etch and two-step self-etch adhesives did not leave any resin residues by wiping and blowing off. All-in-one adhesive revealed resin residues despite wiping off. Prime and Bond NT did not leave any remnant with compomer. Clinicians must be made aware of the consequences of possible adhesive remnants on matrix <span class="hlt">bands</span> that may lead to a defective definitive restoration. The adhesive resin used for Class II restorations may leave resin coats on metal matrix <span class="hlt">bands</span> after polymerization, resulting in <span class="hlt">additional</span> thickness on the metal matrix <span class="hlt">bands</span> and poor quality of the proximal surface of the definitive restoration when the adhesive system is incorporated in the restoration.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/285694-energy-staggering-superdeformed-bands-sup-ce-sup-ce-sup-ce','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/285694-energy-staggering-superdeformed-bands-sup-ce-sup-ce-sup-ce"><span><span class="hlt">Energy</span> staggering in superdeformed <span class="hlt">bands</span> in {sup 131}Ce, {sup 132}Ce, and {sup 133}Ce</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Semple, A.T.; Nolan, P.J.; Beausang, C.W.</p> <p>1996-05-01</p> <p>Superdeformed <span class="hlt">bands</span> observed in {sup 131}Ce, {sup 132}Ce, and {sup 133}Ce have sequences of {gamma}-ray transition <span class="hlt">energies</span> that exhibit a {Delta}{ital I}=2 staggering. This staggering has different characteristics to that seen in previously known cases in other mass regions. The <span class="hlt">energy</span> staggering starts at low rotational frequency ({sq_bullet}{omega}=3 MeV for {sup 131}Ce) at a magnitude of {approximately}{plus_minus}0.3 keV, dies away to zero at intermediate frequency ({sq_bullet}{omega}=0.6{minus}0.7 MeV), and reappears at higher frequencies ({sq_bullet}{omega}{approximately}0.7 MeV). {copyright} {ital 1996 The American Physical Society.}</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27356202','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27356202"><span>Removing <span class="hlt">energy</span> from a beverage influences later food intake more than the same <span class="hlt">energy</span> <span class="hlt">addition</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>McCrickerd, K; Salleh, N B; Forde, C G</p> <p>2016-10-01</p> <p>Designing reduced-calorie foods and beverages without compromising their satiating effect could benefit weight management, assuming that consumers do not compensate for the missing calories at other meals. Though research has demonstrated that compensation for overfeeding is relatively limited, the extent to which <span class="hlt">energy</span> reductions trigger adjustments in later food intake is less clear. The current study tested satiety responses (characterised by changes in appetite and later food intake) to both a covert 200 kcal reduction and an <span class="hlt">addition</span> of maltodextrin to a soymilk test beverage. Twenty-nine healthy male participants were recruited to consume three sensory-matched soymilk beverages across four non-consecutive study days: a medium <span class="hlt">energy</span> control (ME: 300 kcal) and a lower <span class="hlt">energy</span> (LE: 100 kcal) and higher <span class="hlt">energy</span> (HE: 500 kcal) version. The ME control was consumed twice to assess individual consistency in responses to this beverage. Participants were unaware of the <span class="hlt">energy</span> differences across the soymilks. Lunch intake 60 min later increased in response to the LE soymilk, but was unchanged after consuming the HE version. These adjustments accounted for 40% of the <span class="hlt">energy</span> removed from the soymilk and 13% of the <span class="hlt">energy</span> added in. Rated appetite was relatively unaffected by the soymilk <span class="hlt">energy</span> content. No further adjustments were noted for the rest of the day. These data suggest that adult men tested were more sensitive to calorie dilution than calorie <span class="hlt">addition</span> to a familiar beverage. Copyright © 2016 Elsevier Ltd. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014PhyE...59...15B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014PhyE...59...15B"><span>Raman <span class="hlt">bands</span> in Ag nanoparticles obtained in extract of Opuntia ficus-indica plant</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bocarando-Chacon, J.-G.; Cortez-Valadez, M.; Vargas-Vazquez, D.; Rodríguez Melgarejo, F.; Flores-Acosta, M.; Mani-Gonzalez, P. G.; Leon-Sarabia, E.; Navarro-Badilla, A.; Ramírez-Bon, R.</p> <p>2014-05-01</p> <p>Silver nanoparticles have been obtained in an extract of Opuntia ficus-indica plant. The size and distribution of nanoparticles were quantified by atomic force microscopy (AFM). The diameter was estimated to be about 15 nm. In <span class="hlt">addition</span>, <span class="hlt">energy</span> dispersive X-ray spectroscopy (EDX) peaks of silver were observed in these samples. Three Raman <span class="hlt">bands</span> have been experimentally detected at 83, 110 and 160 cm-1. The <span class="hlt">bands</span> at 83 and 110 cm-1 are assigned to the silver-silver Raman modes (skeletal modes) and the Raman mode located at 160 cm-1 has been assigned to breathing modes. Vibrational assignments of Raman modes have been carried out based on the Density Functional Theory (DFT) quantum mechanical calculation. Structural and vibrational properties for small Agn clusters with 2≤n≤9 were determined. Calculated Raman modes for small metal clusters have an approximation trend of Raman <span class="hlt">bands</span>. These Raman <span class="hlt">bands</span> were obtained experimentally for silver nanoparticles (AgNP).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016JAP...119x5701P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016JAP...119x5701P"><span>Ionization equilibrium at the transition from valence-<span class="hlt">band</span> to acceptor-<span class="hlt">band</span> migration of holes in boron-doped diamond</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Poklonski, N. A.; Vyrko, S. A.; Poklonskaya, O. N.; Kovalev, A. I.; Zabrodskii, A. G.</p> <p>2016-06-01</p> <p>A quasi-classical model of ionization equilibrium in the p-type diamond between hydrogen-like acceptors (boron atoms which substitute carbon atoms in the crystal lattice) and holes in the valence <span class="hlt">band</span> (v-<span class="hlt">band</span>) is proposed. The model is applicable on the insulator side of the insulator-metal concentration phase transition (Mott transition) in p-Dia:B crystals. The densities of the spatial distributions of impurity atoms (acceptors and donors) and of holes in the crystal are considered to be Poissonian, and the fluctuations of their electrostatic potential <span class="hlt">energy</span> are considered to be Gaussian. The model accounts for the decrease in thermal ionization <span class="hlt">energy</span> of boron atoms with increasing concentration, as well as for electrostatic fluctuations due to the Coulomb interaction limited to two nearest point charges (impurity ions and holes). The mobility edge of holes in the v-<span class="hlt">band</span> is assumed to be equal to the sum of the threshold <span class="hlt">energy</span> for diffusion percolation and the exchange <span class="hlt">energy</span> of the holes. On the basis of the virial theorem, the temperature Tj is determined, in the vicinity of which the dc <span class="hlt">band</span>-like conductivity of holes in the v-<span class="hlt">band</span> is approximately equal to the hopping conductivity of holes via the boron atoms. For compensation ratio (hydrogen-like donor to acceptor concentration ratio) K ≈ 0.15 and temperature Tj, the concentration of "free" holes in the v-<span class="hlt">band</span> and their jumping (turbulent) drift mobility are calculated. Dependence of the differential <span class="hlt">energy</span> of thermal ionization of boron atoms (at the temperature 3Tj/2) as a function of their concentration N is calculated. The estimates of the extrapolated into the temperature region close to Tj hopping drift mobility of holes hopping from the boron atoms in the charge states (0) to the boron atoms in the charge states (-1) are given. Calculations based on the model show good agreement with electrical conductivity and Hall effect measurements for p-type diamond with boron atom concentrations in the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.dtic.mil/docs/citations/ADA615115','DTIC-ST'); return false;" href="http://www.dtic.mil/docs/citations/ADA615115"><span>First Principles Study of <span class="hlt">Band</span> Structure and <span class="hlt">Band</span> Gap Engineering in Graphene for Device Applications</span></a></p> <p><a target="_blank" href="http://www.dtic.mil/">DTIC Science & Technology</a></p> <p></p> <p>2015-03-20</p> <p>In the bandstructure of graphene which is dominated by Dirac description, valence and conduction <span class="hlt">bands</span> cross the Fermi level at a single point (K...of <span class="hlt">energy</span> <span class="hlt">bands</span> and appearance of Dirac cones near the ‘K’ point and Fermi level the electrons behave like massless Dirac fermions. For applications...results. Introduction Graphene, the super carbon , is now accepted as wonder material with new physics and it has caused major</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015PhRvP...4e4012Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015PhRvP...4e4012Y"><span><span class="hlt">Band</span>-Gap and <span class="hlt">Band</span>-Edge Engineering of Multicomponent Garnet Scintillators from First Principles</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yadav, Satyesh K.; Uberuaga, Blas P.; Nikl, Martin; Jiang, Chao; Stanek, Christopher R.</p> <p>2015-11-01</p> <p>Complex doping schemes in R3 Al5 O12 (where R is the rare-earth element) garnet compounds have recently led to pronounced improvements in scintillator performance. Specifically, by admixing lutetium and yttrium aluminate garnets with gallium and gadolinium, the <span class="hlt">band</span> gap is altered in a manner that facilitates the removal of deleterious electron trapping associated with cation antisite defects. Here, we expand upon this initial work to systematically investigate the effect of substitutional admixing on the <span class="hlt">energy</span> levels of <span class="hlt">band</span> edges. Density-functional theory and hybrid density-functional theory (HDFT) are used to survey potential admixing candidates that modify either the conduction-<span class="hlt">band</span> minimum (CBM) or valence-<span class="hlt">band</span> maximum (VBM). We consider two sets of compositions based on Lu3 B5O12 where B is Al, Ga, In, As, and Sb, and R3Al5 O12 , where R is Lu, Gd, Dy, and Er. We find that admixing with various R cations does not appreciably affect the <span class="hlt">band</span> gap or <span class="hlt">band</span> edges. In contrast, substituting Al with cations of dissimilar ionic radii has a profound impact on the <span class="hlt">band</span> structure. We further show that certain dopants can be used to selectively modify only the CBM or the VBM. Specifically, Ga and In decrease the <span class="hlt">band</span> gap by lowering the CBM, while As and Sb decrease the <span class="hlt">band</span> gap by raising the VBM, the relative change in <span class="hlt">band</span> gap is quantitatively validated by HDFT. These results demonstrate a powerful approach to quickly screen the impact of dopants on the electronic structure of scintillator compounds, identifying those dopants which alter the <span class="hlt">band</span> edges in very specific ways to eliminate both electron and hole traps responsible for performance limitations. This approach should be broadly applicable for the optimization of electronic and optical performance for a wide range of compounds by tuning the VBM and CBM.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2011-title47-vol1/pdf/CFR-2011-title47-vol1-sec15-713.pdf','CFR2011'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2011-title47-vol1/pdf/CFR-2011-title47-vol1-sec15-713.pdf"><span>47 CFR 15.713 - TV <span class="hlt">bands</span> database.</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collectionCfr.action?selectedYearFrom=2011&page.go=Go">Code of Federal Regulations, 2011 CFR</a></p> <p></p> <p>2011-10-01</p> <p>... authorized services operating in the TV <span class="hlt">bands</span>. In <span class="hlt">addition</span>, a TV <span class="hlt">bands</span> database must also verify that the FCC identifier (FCC ID) of a device seeking access to its services is valid; under this requirement the TV <span class="hlt">bands</span>... information will come from the official Commission database. These services include: (i) Digital television...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28799383','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28799383"><span>Focus: Nucleation kinetics of shear <span class="hlt">bands</span> in metallic glass.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wang, J Q; Perepezko, J H</p> <p>2016-12-07</p> <p>The development of shear <span class="hlt">bands</span> is recognized as the primary mechanism in controlling the plastic deformability of metallic glasses. However, the kinetics of the nucleation of shear <span class="hlt">bands</span> has received limited attention. The nucleation of shear <span class="hlt">bands</span> in metallic glasses (MG) can be investigated using a nanoindentation method to monitor the development of the first pop-in event that is a signature of shear <span class="hlt">band</span> nucleation. The analysis of a statistically significant number of first pop-in events demonstrates the stochastic behavior that is characteristic of nucleation and reveals a multimodal behavior associated with local spatial heterogeneities. The shear <span class="hlt">band</span> nucleation rate of the two nucleation modes and the associated activation <span class="hlt">energy</span>, activation volume, and site density were determined by loading rate experiments. The nucleation activation <span class="hlt">energy</span> is very close to the value that is characteristic of the β relaxation in metallic glass. The identification of the rate controlling kinetics for shear <span class="hlt">band</span> nucleation offers guidance for promoting plastic flow in metallic glass.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29488385','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29488385"><span>Origin of the Two <span class="hlt">Bands</span> in the B800 Ring and Their Involvement in the <span class="hlt">Energy</span> Transfer Network of Allochromatium vinosum.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Schröter, Marco; Alcocer, Marcelo J P; Cogdell, Richard J; Kühn, Oliver; Zigmantas, Donatas</p> <p>2018-03-15</p> <p>Bacterial photosynthesis features robust and adaptable <span class="hlt">energy</span>-harvesting processes in which light-harvesting proteins play a crucial role. The peripheral light-harvesting complex of the purple bacterium Allochromatium vinosum is particularly distinct, featuring a double peak structure in its B800 absorption <span class="hlt">band</span>. Two hypotheses-not necessarily mutually exclusive-concerning the origin of this splitting have been proposed; either two distinct B800 bacteriochlorophyll site <span class="hlt">energies</span> are involved, or an excitonic dimerization of bacteriochlorophylls within the B800 ring takes place. Through the use of two-dimensional electronic spectroscopy, we present unambiguous evidence that excitonic interaction shapes the split <span class="hlt">band</span>. We further identify and characterize all of the <span class="hlt">energy</span> transfer pathways within this complex by using a global kinetic fitting procedure. Our approach demonstrates how the combination of two-dimensional spectral resolution and self-consistent fitting allows for extraction of information on light-harvesting processes, which would otherwise be inaccessible due to signal congestion.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012PhRvC..86d4305W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012PhRvC..86d4305W"><span>Rotational <span class="hlt">band</span> properties of 173W</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wang, H. X.; Zhang, Y. H.; Zhou, X. H.; Liu, M. L.; Ding, B.; Li, G. S.; Hua, W.; Zhou, H. B.; Guo, S.; Qiang, Y. H.; Oshima, M.; Koizumi, M.; Toh, Y.; Kimura, A.; Harada, H.; Furutaka, K.; Kitatani, F.; Nakamura, S.; Hatsukawa, Y.; Ohta, M.; Hara, K.; Kin, T.; Meng, J.</p> <p>2012-10-01</p> <p>High-spin states in 173W have been studied using the 150Nd(28Si,5n)173W reaction at beam <span class="hlt">energies</span> of 135 and 140 MeV. The previously known <span class="hlt">bands</span> associated with the 7/2+[633], 5/2-[512], and 1/2-[521] configurations are extended significantly, and the unfavored signature branch of the 1/2-[521] <span class="hlt">band</span> is established for the first time. The <span class="hlt">band</span> properties, such as level spacings, <span class="hlt">band</span>-crossing frequencies, alignment gains, and signature splittings, are discussed with an emphasis on the low-spin signature inversion observed in the 5/2-[512] <span class="hlt">band</span>. By comparing the experimental B(M1)/B(E2) ratios with the theoretical values, we conclude that the configuration of the 5/2-[512] <span class="hlt">band</span> is quite pure at low spins without appreciable admixture of the 5/2-[523] orbit, in conflict with the particle rotor model calculated results.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19950009415','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19950009415"><span>Diffuse interstellar <span class="hlt">bands</span> in reflection nebulae</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Fischer, O.; Henning, Thomas; Pfau, Werner; Stognienko, R.</p> <p>1994-01-01</p> <p>A Monte Carlo code for radiation transport calculations is used to compare the profiles of the lambda lambda 5780 and 6613 Angstrom diffuse interstellar <span class="hlt">bands</span> in the transmitted and the reflected light of a star embedded within an optically thin dust cloud. In <span class="hlt">addition</span>, the behavior of polarization across the <span class="hlt">bands</span> were calculated. The wavelength dependent complex indices of refraction across the <span class="hlt">bands</span> were derived from the embedded cavity model. In view of the existence of different families of diffuse interstellar <span class="hlt">bands</span> the question of other parameters of influence is addressed in short.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19870014404','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19870014404"><span>Block 3 X-<span class="hlt">band</span> receiver-exciter</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Johns, C. E.</p> <p>1987-01-01</p> <p>The development of an X-<span class="hlt">band</span> exciter, for use in the X-<span class="hlt">Band</span> Uplink Subsystem, was completed. The exciter generates the drive signal for the X-<span class="hlt">band</span> transmitter and also generates coherent test signals for the S- and X-<span class="hlt">band</span> Block 3 translator and a Doppler reference signal for the Doppler extractor system. In <span class="hlt">addition</span> to the above, the exciter generates other reference signals that are described. Also presented is an overview of the exciter design and some test data taken on the prototype. A brief discussion of the Block 3 Doppler extractor is presented.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017MS%26E..258a2007N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017MS%26E..258a2007N"><span>Microstructure characterisation of Ti-6Al-4V from different <span class="hlt">additive</span> manufacturing processes</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Neikter, M.; Åkerfeldt, P.; Pederson, R.; Antti, M.-L.</p> <p>2017-10-01</p> <p>The focus of this work has been microstructure characterisation of Ti-6Al-4V manufactured by five different <span class="hlt">additive</span> manufacturing (AM) processes. The microstructure features being characterised are the prior β size, grain boundary α and α lath thickness. It was found that material manufactured with powder bed fusion processes has smaller prior β grains than the material from directed <span class="hlt">energy</span> deposition processes. The AM processes with fast cooling rate render in thinner α laths and also thinner, and in some cases discontinuous, grain boundary α. Furthermore, it has been observed that material manufactured with the directed <span class="hlt">energy</span> deposition processes has parallel <span class="hlt">bands</span>, except for one condition when the parameters were changed, while the powder bed fusion processes do not have any parallel <span class="hlt">bands</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018EPJB...91..108P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018EPJB...91..108P"><span>Density-functional <span class="hlt">energy</span> gaps of solids demystified</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Perdew, John P.; Ruzsinszky, Adrienn</p> <p>2018-06-01</p> <p>The fundamental <span class="hlt">energy</span> gap of a solid is a ground-state second <span class="hlt">energy</span> difference. Can one find the fundamental gap from the gap in the <span class="hlt">band</span> structure of Kohn-Sham density functional theory? An argument of Williams and von Barth (WB), 1983, suggests that one can. In fact, self-consistent <span class="hlt">band</span>-structure calculations within the local density approximation or the generalized gradient approximation (GGA) yield the fundamental gap within the same approximation for the <span class="hlt">energy</span>. Such a calculation with the exact density functional would yield a <span class="hlt">band</span> gap that also underestimates the fundamental gap, because the exact Kohn-Sham potential in a solid jumps up by an <span class="hlt">additive</span> constant when one electron is added, and the WB argument does not take this effect into account. The WB argument has been extended recently to generalized Kohn-Sham theory, the simplest way to implement meta-GGAs and hybrid functionals self-consistently, with an exchange-correlation potential that is a non-multiplication operator. Since this operator is continuous, the <span class="hlt">band</span> gap is again the fundamental gap within the same approximation, but, because the approximations are more realistic, so is the <span class="hlt">band</span> gap. What approximations might be even more realistic?</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29401928','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29401928"><span>Efficient full-spectrum utilization, reception and conversion of solar <span class="hlt">energy</span> by broad-<span class="hlt">band</span> nanospiral antenna.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zhao, Huaqiao; Gao, Huotao; Cao, Ting; Li, Boya</p> <p>2018-01-22</p> <p>In this work, the collection of solar <span class="hlt">energy</span> by a broad-<span class="hlt">band</span> nanospiral antenna is investigated in order to solve the low efficiency of the solar rectenna based on conventional nanoantennas. The antenna impedance, radiation, polarization and effective area are all considered in the efficiency calculation using the finite integral technique. The wavelength range investigated is 300-3000 nm, which corresponds to more than 98% of the solar radiation <span class="hlt">energy</span>. It's found that the nanospiral has stronger field enhancement in the gap than a nanodipole counterpart. And a maximum harvesting efficiency about 80% is possible in principle for the nanospiral coupled to a rectifier resistance of 200 Ω, while about 10% for the nanodipole under the same conditions. Moreover, the nanospiral could be coupled to a rectifier diode of high resistance more easily than the nanodipole. These results indicate that the efficient full-spectrum utilization, reception and conversion of solar <span class="hlt">energy</span> can be achieved by the nanospiral antenna, which is expected to promote the solar rectenna to be a promising technology in the clean, renewable <span class="hlt">energy</span> application.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29863833','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29863833"><span><span class="hlt">Band-to-Band</span> Tunneling-Dominated Thermo-Enhanced Field Electron Emission from p-Si/ZnO Nanoemitters.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Huang, Zhizhen; Huang, Yifeng; Xu, Ningsheng; Chen, Jun; She, Juncong; Deng, Shaozhi</p> <p>2018-06-13</p> <p>Thermo-enhancement is an effective way to achieve high performance field electron emitters, and enables the individually tuning on the emission current by temperature and the electron <span class="hlt">energy</span> by voltage. The field emission current from metal or n-doped semiconductor emitter at a relatively lower temperature (i.e., < 1000 K) is less temperature sensitive due to the weak dependence of free electron density on temperature, while that from p-doped semiconductor emitter is restricted by its limited free electron density. Here, we developed full array of uniform individual p-Si/ZnO nanoemitters and demonstrated the strong thermo-enhanced field emission. The mechanism of forming uniform nanoemitters with well Si/ZnO mechanical joint in the nanotemplates was elucidated. No current saturation was observed in the thermo-enhanced field emission measurements. The emission current density showed about ten-time enhancement (from 1.31 to 12.11 mA/cm 2 at 60.6 MV/m) by increasing the temperature from 323 to 623 K. The distinctive performance did not agree with the interband excitation mechanism but well-fit to the <span class="hlt">band-to-band</span> tunneling model. The strong thermo-enhancement was proposed to be benefit from the increase of <span class="hlt">band-to-band</span> tunneling probability at the surface portion of the p-Si/ZnO nanojunction. This work provides promising cathode for portable X-ray tubes/panel, ionization vacuum gauges and low <span class="hlt">energy</span> electron beam lithography, in where electron-dose control at a fixed <span class="hlt">energy</span> is needed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2001nspc.book..249K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2001nspc.book..249K"><span><roman>B</roman>(<roman>M</roman>1) values in the <span class="hlt">band</span>-crossing of shears <span class="hlt">bands</span> in 197<roman>Pb</roman></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Krücken, R.; Cooper, J. R.; Beausang, C. W.; Novak, J. R.; Dewald, A.; Klug, T.; Kemper, G.; von Brentano, P.; Carpenter, M.; Wiedenhöver, I.</p> <p></p> <p>We present details of the <span class="hlt">band</span> crossing mechanism of shears <span class="hlt">bands</span> using the example of 197Pb. Absolute reduced matrix elements B(M1) were determined by means of a RDM lifetime measurement in one of the shears <span class="hlt">bands</span> in 197Pb. The experiment was performed using the New Yale Plunger Device (NYPD) in conjunction with the Gammasphere array. <span class="hlt">Band</span> mixing calculations on the basis of the semi-classical model of the shears mechanism are used to describe the transition matrix elements B(M1) and <span class="hlt">energies</span> throughout the <span class="hlt">band</span>-crossing regions. Good agreement with the data was obtained and the detailed composition of the states in the shears <span class="hlt">band</span> are discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015PhRvL.115f6403C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015PhRvL.115f6403C"><span>Linear Scaling of the Exciton Binding <span class="hlt">Energy</span> versus the <span class="hlt">Band</span> Gap of Two-Dimensional Materials</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Choi, Jin-Ho; Cui, Ping; Lan, Haiping; Zhang, Zhenyu</p> <p>2015-08-01</p> <p>The exciton is one of the most crucial physical entities in the performance of optoelectronic and photonic devices, and widely varying exciton binding <span class="hlt">energies</span> have been reported in different classes of materials. Using first-principles calculations within the G W -Bethe-Salpeter equation approach, here we investigate the excitonic properties of two recently discovered layered materials: phosphorene and graphene fluoride. We first confirm large exciton binding <span class="hlt">energies</span> of, respectively, 0.85 and 2.03 eV in these systems. Next, by comparing these systems with several other representative two-dimensional materials, we discover a striking linear relationship between the exciton binding <span class="hlt">energy</span> and the <span class="hlt">band</span> gap and interpret the existence of the linear scaling law within a simple hydrogenic picture. The broad applicability of this novel scaling law is further demonstrated by using strained graphene fluoride. These findings are expected to stimulate related studies in higher and lower dimensions, potentially resulting in a deeper understanding of excitonic effects in materials of all dimensionalities.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li class="active"><span>11</span></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_11 --> <div id="page_12" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li class="active"><span>12</span></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="221"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/17280325','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/17280325"><span>Production of photocurrent due to intermediate-to-conduction-<span class="hlt">band</span> transitions: a demonstration of a key operating principle of the intermediate-<span class="hlt">band</span> solar cell.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Martí, A; Antolín, E; Stanley, C R; Farmer, C D; López, N; Díaz, P; Cánovas, E; Linares, P G; Luque, A</p> <p>2006-12-15</p> <p>We present intermediate-<span class="hlt">band</span> solar cells manufactured using quantum dot technology that show for the first time the production of photocurrent when two sub-<span class="hlt">band</span>-gap <span class="hlt">energy</span> photons are absorbed simultaneously. One photon produces an optical transition from the intermediate-<span class="hlt">band</span> to the conduction <span class="hlt">band</span> while the second pumps an electron from the valence <span class="hlt">band</span> to the intermediate-<span class="hlt">band</span>. The detection of this two-photon absorption process is essential to verify the principles of operation of the intermediate-<span class="hlt">band</span> solar cell. The phenomenon is the cornerstone physical principle that ultimately allows the production of photocurrent in a solar cell by below <span class="hlt">band</span> gap photon absorption, without degradation of its output voltage.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018NJPh...20c3005M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018NJPh...20c3005M"><span>Non-<span class="hlt">additive</span> dissipation in open quantum networks out of equilibrium</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mitchison, Mark T.; Plenio, Martin B.</p> <p>2018-03-01</p> <p>We theoretically study a simple non-equilibrium quantum network whose dynamics can be expressed and exactly solved in terms of a time-local master equation. Specifically, we consider a pair of coupled fermionic modes, each one locally exchanging <span class="hlt">energy</span> and particles with an independent, macroscopic thermal reservoir. We show that the generator of the asymptotic master equation is not <span class="hlt">additive</span>, i.e. it cannot be expressed as a sum of contributions describing the action of each reservoir alone. Instead, we identify an <span class="hlt">additional</span> interference term that generates coherences in the <span class="hlt">energy</span> eigenbasis, associated with the current of conserved particles flowing in the steady state. Notably, non-<span class="hlt">additivity</span> arises even for wide-<span class="hlt">band</span> reservoirs coupled arbitrarily weakly to the system. Our results shed light on the non-trivial interplay between multiple thermal noise sources in modular open quantum systems.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23244696','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23244696"><span>Effects of adsorbed pyridine derivatives and ultrathin atomic-layer-deposited alumina coatings on the conduction <span class="hlt">band</span>-edge <span class="hlt">energy</span> of TiO2 and on redox-shuttle-derived dark currents.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Katz, Michael J; Vermeer, Michael J D; Farha, Omar K; Pellin, Michael J; Hupp, Joseph T</p> <p>2013-01-15</p> <p>Both the adsorption of t-butylpyridine and the atomic-layer deposition of ultrathin conformal coatings of insulators (such as alumina) are known to boost open-circuit photovoltages substantially for dye-sensitized solar cells. One attractive interpretation is that these modifiers significantly shift the conduction-edge <span class="hlt">energy</span> of the electrode, thereby shifting the onset potential for dark current arising from the interception of injected electrons by solution-phase redox shuttle components such as Co(phenanthroline)(3)(3+) and triiodide. For standard, high-area, nanoporous photoelectrodes, <span class="hlt">band</span>-edge <span class="hlt">energies</span> are difficult to measure directly. In contrast, for flat electrodes they are readily accessible from Mott-Schottky analyses of impedance data. Using such electrodes (specifically TiO(2)), we find that neither organic nor inorganic electrode-surface modifiers shift the conduction-<span class="hlt">band</span>-edge <span class="hlt">energy</span> sufficiently to account fully for the beneficial effects on electrode behavior (i.e., the suppression of dark current). <span class="hlt">Additional</span> experiments reveal that the efficacy of ultrathin coatings of Al(2)O(3) arises chiefly from the passivation of redox-catalytic surface states. In contrast, adsorbed t-butylpyridine appears to suppress dark currents mainly by physically blocking access of shuttle molecules to the electrode surface. Studies with other derivatives of pyridine, including sterically and/or electronically diverse derivatives, show that heterocycle adsorption and the concomitant suppression of dark current does not require the coordination of surface Ti(IV) or Al(III) atoms. Notably, the favorable (i.e., negative) shifts in onset potential for the flow of dark current engendered by organic and inorganic surface modifiers are <span class="hlt">additive</span>. Furthermore, they appear to be largely insensitive to the identity of shuttle molecules.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016JMoSp.322...18S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016JMoSp.322...18S"><span>Near UV <span class="hlt">bands</span> of jet-cooled CaO</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Stewart, Jacob T.; Sullivan, Michael N.; Heaven, Michael C.</p> <p>2016-04-01</p> <p>The electronic spectrum of CaO has been recorded for the 29,800-33,150 cm-1 <span class="hlt">energy</span> range. Jet cooling was used to obtain relatively uncongested spectra. Rotationally resolved <span class="hlt">bands</span> have been assigned to the C1Σ+-X1Σ+ and F1∏-X transitions. These data extend the range of vibronic levels characterized for the upper states. Three <span class="hlt">additional</span> vibronic states were observed as a short progression. One of these levels, which are of 0+ symmetry, interacts strongly with the C1Σ+, v‧ = 7 level. Possible assignments for the perturbing state are considered.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018ApSS..433..530Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018ApSS..433..530Y"><span><span class="hlt">Energy-band</span> alignment of (HfO2)x(Al2O3)1-x gate dielectrics deposited by atomic layer deposition on β-Ga2O3 (-201)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yuan, Lei; Zhang, Hongpeng; Jia, Renxu; Guo, Lixin; Zhang, Yimen; Zhang, Yuming</p> <p>2018-03-01</p> <p><span class="hlt">Energy</span> <span class="hlt">band</span> alignments between series <span class="hlt">band</span> of Al-rich high-k materials (HfO2)x(Al2O3)1-x and β-Ga2O3 are investigated using X-Ray Photoelectron Spectroscopy (XPS). The results exhibit sufficient conduction <span class="hlt">band</span> offsets (1.42-1.53 eV) in (HfO2)x(Al2O3)1-x/β-Ga2O3. In <span class="hlt">addition</span>, it is also obtained that the value of Eg, △Ec, and △Ev for (HfO2)x(Al2O3)1-x/β-Ga2O3 change linearly with x, which can be expressed by 6.98-1.27x, 1.65-0.56x, and 0.48-0.70x, respectively. The higher dielectric constant and higher effective breakdown electric field of (HfO2)x(Al2O3)1-x compared with Al2O3, coupled with sufficient barrier height and lower gate leakage makes it a potential dielectric for high voltage β-Ga2O3 power MOSFET, and also provokes interest in further investigation of HfAlO/β-Ga2O3 interface properties.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017ApSS..420..523C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017ApSS..420..523C"><span>In-situ growth of HfO2 on clean 2H-MoS2 surface: Growth mode, interface reactions and <span class="hlt">energy</span> <span class="hlt">band</span> alignment</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Chen, Chang Pang; Ong, Bin Leong; Ong, Sheau Wei; Ong, Weijie; Tan, Hui Ru; Chai, Jian Wei; Zhang, Zheng; Wang, Shi Jie; Pan, Ji Sheng; Harrison, Leslie John; Kang, Hway Chuan; Tok, Eng Soon</p> <p>2017-10-01</p> <p>Room temperature growth of HfO2 thin film on clean 2H-MoS2 via plasma-sputtering of Hf-metal target in an argon/oxygen environment was studied in-situ using x-ray photoelectron spectroscopy (XPS). The deposited film was observed to grow akin to a layer-by-layer growth mode. At the onset of growth, a mixture of sulfate- and sulfite-like species (SOx2- where x = 3, 4), and molybdenum trioxide (MoO3), are formed at the HfO2/MoS2 interface. An initial decrease in binding <span class="hlt">energies</span> for both Mo 3d and S 2p core-levels of the MoS2 substrate by 0.4 eV was also observed. Their binding <span class="hlt">energies</span>, however, did not change further with increasing HfO2 thickness. There was no observable change in the Hf4f core-level binding <span class="hlt">energy</span> throughout the deposition process. With increasing HfO2 deposition, MoO3 becomes buried at the interface while SOx2- was observed to be present in the film. The shift of 0.4 eV for both Mo 3d and S 2p core-levels of the MoS2 substrate can be attributed to a charge transfer from the substrate to the MoO3/SOx2--like interface layer. Consequently, the Type I heterojunction valence <span class="hlt">band</span> offset (conduction <span class="hlt">band</span> offset) becomes 1.7 eV (2.9 eV) instead of 1.3 eV (3.3 eV) expected from considering the bulk HfO2 and MoS2 valence <span class="hlt">band</span> offset (conduction <span class="hlt">band</span> offset). The formation of these states and its influence on <span class="hlt">band</span> offsets will need to be considered in their device applications.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23556883','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23556883"><span>Feasibility of producing a short, high <span class="hlt">energy</span> s-<span class="hlt">band</span> linear accelerator using a klystron power source.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Baillie, Devin; St Aubin, J; Fallone, B G; Steciw, S</p> <p>2013-04-01</p> <p>To use a finite-element method (FEM) model to study the feasibility of producing a short s-<span class="hlt">band</span> (2.9985 GHz) waveguide capable of producing x-rays <span class="hlt">energies</span> up to 10 MV, for applications in a linac-MR, as well as conventional radiotherapy. An existing waveguide FEM model developed by the authors' group is used to simulate replacing the magnetron power source with a klystron. Peak fields within the waveguide are compared with a published experimental threshold for electric breakdown. The RF fields in the first accelerating cavity are scaled, approximating the effect of modifications to the first coupling cavity. Electron trajectories are calculated within the RF fields, and the <span class="hlt">energy</span> spectrum, beam current, and focal spot of the electron beam are analyzed. One electron spectrum is selected for Monte Carlo simulations and the resulting PDD compared to measurement. When the first cavity fields are scaled by a factor of 0.475, the peak magnitude of the electric fields within the waveguide are calculated to be 223.1 MV∕m, 29% lower than the published threshold for breakdown at this operating frequency. Maximum electron <span class="hlt">energy</span> increased from 6.2 to 10.4 MeV, and beam current increased from 134 to 170 mA. The focal spot FWHM is decreased slightly from 0.07 to 0.05 mm, and the width of the <span class="hlt">energy</span> spectrum increased slightly from 0.44 to 0.70 MeV. Monte Carlo results show dmax is at 2.15 cm for a 10 × 10 cm(2) field, compared with 2.3 cm for a Varian 10 MV linac, while the penumbral widths are 4.8 and 5.6 mm, respectively. The authors' simulation results show that a short, high-<span class="hlt">energy</span>, s-<span class="hlt">band</span> accelerator is feasible and electric breakdown is not expected to interfere with operation at these field strengths. With minor modifications to the first coupling cavity, all electron beam parameters are improved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016EPJAP..7320301Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016EPJAP..7320301Y"><span><span class="hlt">Energy</span> <span class="hlt">band</span> and transport properties in magnetic aperiodic graphene superlattices of Thue-Morse sequence</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yin, Yiheng; Niu, Yanxiong; Zhang, Huiyun; Zhang, Yuping; Liu, Haiyue</p> <p>2016-02-01</p> <p>Utilizing the transfer matrix method, we develop the electronic <span class="hlt">band</span> structure and transport properties in Thue-Morse aperiodic graphene superlattices with magnetic barriers. It is found that the normal transmission is blocked and the position of the Dirac point can be shifted along the wavevector axis by changing the height and width ratio of magnetic barriers, which is intrinsic different from electronic field modulated superlattices. In <span class="hlt">addition</span>, the angular threshold property of the transmission spectra and the oscillatory property of the conductance have been studied.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016PhRvB..94c5201R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016PhRvB..94c5201R"><span>Zero-phonon line and fine structure of the yellow luminescence <span class="hlt">band</span> in GaN</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Reshchikov, M. A.; McNamara, J. D.; Zhang, F.; Monavarian, M.; Usikov, A.; Helava, H.; Makarov, Yu.; Morkoç, H.</p> <p>2016-07-01</p> <p>The yellow luminescence <span class="hlt">band</span> was studied in undoped and Si-doped GaN samples by steady-state and time-resolved photoluminescence. At low temperature (18 K), the zero-phonon line (ZPL) for the yellow <span class="hlt">band</span> is observed at 2.57 eV and attributed to electron transitions from a shallow donor to a deep-level defect. At higher temperatures, the ZPL at 2.59 eV emerges, which is attributed to electron transitions from the conduction <span class="hlt">band</span> to the same defect. In <span class="hlt">addition</span> to the ZPL, a set of phonon replicas is observed, which is caused by the emission of phonons with <span class="hlt">energies</span> of 39.5 meV and 91.5 meV. The defect is called the YL1 center. The possible identity of the YL1 center is discussed. The results indicate that the same defect is responsible for the strong YL1 <span class="hlt">band</span> in undoped and Si-doped GaN samples.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/18764346','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/18764346"><span>Transition-metal-substituted indium thiospinels as novel intermediate-<span class="hlt">band</span> materials: prediction and understanding of their electronic properties.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Palacios, P; Aguilera, I; Sánchez, K; Conesa, J C; Wahnón, P</p> <p>2008-07-25</p> <p>Results of density-functional calculations for indium thiospinel semiconductors substituted at octahedral sites with isolated transition metals (M=Ti,V) show an isolated partially filled narrow <span class="hlt">band</span> containing three t2g-type states per M atom inside the usual semiconductor <span class="hlt">band</span> gap. Thanks to this electronic structure feature, these materials will allow the absorption of photons with <span class="hlt">energy</span> below the <span class="hlt">band</span> gap, in <span class="hlt">addition</span> to the normal light absorption of a semiconductor. To our knowledge, we demonstrate for the first time the formation of an isolated intermediate electronic <span class="hlt">band</span> structure through M substitution at octahedral sites in a semiconductor, leading to an enhancement of the absorption coefficient in both infrared and visible ranges of the solar spectrum. This electronic structure feature could be applied for developing a new third-generation photovoltaic cell.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JEMat.tmp.2704M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JEMat.tmp.2704M"><span>Tuning the <span class="hlt">Energy</span> Gap of SiCH3 Nanomaterials Under Elastic Strain</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ma, Shengqian; Li, Feng; Geng, Jiguo; Zhu, Mei; Li, Suyan; Han, Juguang</p> <p>2018-05-01</p> <p>SiCH3 nanomaterials have been studied using the density functional theory. When the nanosheets and nanoribbons (armchair and zigzag) are introduced, their <span class="hlt">energy</span> gap is modulated under elastic strain and width. The results show that the <span class="hlt">band</span> gap of SiCH3 nanomaterials can be easily tuned using elastic strains and widths. Surprisingly, the <span class="hlt">band</span> gap can be modulated along two directions, namely, compressing and stretching. The <span class="hlt">band</span> gap decreases when increasing stretching strain or decreasing compressing strain. In <span class="hlt">addition</span>, the <span class="hlt">band</span> gap decreases when increasing the nanoribbon width. For <span class="hlt">energy</span> gap engineering, the <span class="hlt">band</span> gap can be tuned by strains and widths. Therefore, the SiCH3 nanomaterials play important roles in potential applications for strain sensors, electronics, and optical electronics.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24265209','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24265209"><span>Significant reduction in NiO <span class="hlt">band</span> gap upon formation of Lix Ni1-x O alloys: applications to solar <span class="hlt">energy</span> conversion.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Alidoust, Nima; Toroker, Maytal Caspary; Keith, John A; Carter, Emily A</p> <p>2014-01-01</p> <p>Long-term sustainable solar <span class="hlt">energy</span> conversion relies on identifying economical and versatile semiconductor materials with appropriate <span class="hlt">band</span> structures for photovoltaic and photocatalytic applications (e.g., <span class="hlt">band</span> gaps of ∼ 1.5-2.0 eV). Nickel oxide (NiO) is an inexpensive yet highly promising candidate. Its charge-transfer character may lead to longer carrier lifetimes needed for higher efficiencies, and its conduction <span class="hlt">band</span> edge is suitable for driving hydrogen evolution via water-splitting. However, NiO's large <span class="hlt">band</span> gap (∼ 4 eV) severely limits its use in practical applications. Our first-principles quantum mechanics calculations show <span class="hlt">band</span> gaps dramatically decrease to ∼ 2.0 eV when NiO is alloyed with Li2O. We show that Lix Ni1-x O alloys (with x=0.125 and 0.25) are p-type semiconductors, contain states with no impurity levels in the gap and maintain NiO's desirable charge-transfer character. Lastly, we show that the alloys have potential for photoelectrochemical applications, with <span class="hlt">band</span> edges well-placed for photocatalytic hydrogen production and CO2 reduction, as well as in tandem dye-sensitized solar cells as a photocathode. Copyright © 2014 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016JAP...120d4307H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016JAP...120d4307H"><span>Modeling of multi-<span class="hlt">band</span> drift in nanowires using a full <span class="hlt">band</span> Monte Carlo simulation</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hathwar, Raghuraj; Saraniti, Marco; Goodnick, Stephen M.</p> <p>2016-07-01</p> <p>We report on a new numerical approach for multi-<span class="hlt">band</span> drift within the context of full <span class="hlt">band</span> Monte Carlo (FBMC) simulation and apply this to Si and InAs nanowires. The approach is based on the solution of the Krieger and Iafrate (KI) equations [J. B. Krieger and G. J. Iafrate, Phys. Rev. B 33, 5494 (1986)], which gives the probability of carriers undergoing interband transitions subject to an applied electric field. The KI equations are based on the solution of the time-dependent Schrödinger equation, and previous solutions of these equations have used Runge-Kutta (RK) methods to numerically solve the KI equations. This approach made the solution of the KI equations numerically expensive and was therefore only applied to a small part of the Brillouin zone (BZ). Here we discuss an alternate approach to the solution of the KI equations using the Magnus expansion (also known as "exponential perturbation theory"). This method is more accurate than the RK method as the solution lies on the exponential map and shares important qualitative properties with the exact solution such as the preservation of the unitary character of the time evolution operator. The solution of the KI equations is then incorporated through a modified FBMC free-flight drift routine and applied throughout the nanowire BZ. The importance of the multi-<span class="hlt">band</span> drift model is then demonstrated for the case of Si and InAs nanowires by simulating a uniform field FBMC and analyzing the average carrier <span class="hlt">energies</span> and carrier populations under high electric fields. Numerical simulations show that the average <span class="hlt">energy</span> of the carriers under high electric field is significantly higher when multi-<span class="hlt">band</span> drift is taken into consideration, due to the interband transitions allowing carriers to achieve higher <span class="hlt">energies</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018PhRvC..97b4308H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PhRvC..97b4308H"><span>Multiple <span class="hlt">band</span> structures in 70Ge</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Haring-Kaye, R. A.; Morrow, S. I.; Döring, J.; Tabor, S. L.; Le, K. Q.; Allegro, P. R. P.; Bender, P. C.; Elder, R. M.; Medina, N. H.; Oliveira, J. R. B.; Tripathi, Vandana</p> <p>2018-02-01</p> <p>High-spin states in 70Ge were studied using the 55Mn(18O,p 2 n ) fusion-evaporation reaction at a beam <span class="hlt">energy</span> of 50 MeV. Prompt γ -γ coincidences were measured using the Florida State University Compton-suppressed Ge array consisting of three Clover detectors and seven single-crystal detectors. An investigation of these coincidences resulted in the <span class="hlt">addition</span> of 31 new transitions and the rearrangement of four others in the 70Ge level scheme, providing a more complete picture of the high-spin decay pattern involving both positive- and negative-parity states with multiple <span class="hlt">band</span> structures. Spins were assigned based on directional correlation of oriented nuclei ratios, which many times also led to unambiguous parity determinations based on the firm assignments for low-lying states made in previous work. Total Routhian surface calculations, along with the observed trends in the experimental kinematic moment of inertia with rotational frequency, support the multiquasiparticle configurations of the various crossing <span class="hlt">bands</span> proposed in recent studies. The high-spin excitation spectra predicted by previous shell-model calculations compare favorably with the experimental one determined from this study.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22596831-ionization-equilibrium-transition-from-valence-band-acceptor-band-migration-holes-boron-doped-diamond','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22596831-ionization-equilibrium-transition-from-valence-band-acceptor-band-migration-holes-boron-doped-diamond"><span>Ionization equilibrium at the transition from valence-<span class="hlt">band</span> to acceptor-<span class="hlt">band</span> migration of holes in boron-doped diamond</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Poklonski, N. A., E-mail: poklonski@bsu.by; Vyrko, S. A.; Poklonskaya, O. N.</p> <p></p> <p>A quasi-classical model of ionization equilibrium in the p-type diamond between hydrogen-like acceptors (boron atoms which substitute carbon atoms in the crystal lattice) and holes in the valence <span class="hlt">band</span> (v-<span class="hlt">band</span>) is proposed. The model is applicable on the insulator side of the insulator–metal concentration phase transition (Mott transition) in p-Dia:B crystals. The densities of the spatial distributions of impurity atoms (acceptors and donors) and of holes in the crystal are considered to be Poissonian, and the fluctuations of their electrostatic potential <span class="hlt">energy</span> are considered to be Gaussian. The model accounts for the decrease in thermal ionization <span class="hlt">energy</span> of boron atomsmore » with increasing concentration, as well as for electrostatic fluctuations due to the Coulomb interaction limited to two nearest point charges (impurity ions and holes). The mobility edge of holes in the v-<span class="hlt">band</span> is assumed to be equal to the sum of the threshold <span class="hlt">energy</span> for diffusion percolation and the exchange <span class="hlt">energy</span> of the holes. On the basis of the virial theorem, the temperature T{sub j} is determined, in the vicinity of which the dc <span class="hlt">band</span>-like conductivity of holes in the v-<span class="hlt">band</span> is approximately equal to the hopping conductivity of holes via the boron atoms. For compensation ratio (hydrogen-like donor to acceptor concentration ratio) K ≈ 0.15 and temperature T{sub j}, the concentration of “free” holes in the v-<span class="hlt">band</span> and their jumping (turbulent) drift mobility are calculated. Dependence of the differential <span class="hlt">energy</span> of thermal ionization of boron atoms (at the temperature 3T{sub j}/2) as a function of their concentration N is calculated. The estimates of the extrapolated into the temperature region close to T{sub j} hopping drift mobility of holes hopping from the boron atoms in the charge states (0) to the boron atoms in the charge states (−1) are given. Calculations based on the model show good agreement with electrical conductivity and Hall effect measurements for p</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1248741-band-gap-band-edge-engineering-multicomponent-garnet-scintillators-from-first-principles','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1248741-band-gap-band-edge-engineering-multicomponent-garnet-scintillators-from-first-principles"><span><span class="hlt">Band</span>-gap and <span class="hlt">band</span>-edge engineering of multicomponent garnet scintillators from first principles</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Yadav, Satyesh K.; Uberuaga, Blas P.; Nikl, Martin; ...</p> <p>2015-11-24</p> <p>Complex doping schemes in R 3Al 5O 12 (where R is the rare-earth element) garnet compounds have recently led to pronounced improvements in scintillator performance. Specifically, by admixing lutetium and yttrium aluminate garnets with gallium and gadolinium, the <span class="hlt">band</span> gap is altered in a manner that facilitates the removal of deleterious electron trapping associated with cation antisite defects. Here, we expand upon this initial work to systematically investigate the effect of substitutional admixing on the <span class="hlt">energy</span> levels of <span class="hlt">band</span> edges. Density-functional theory and hybrid density-functional theory (HDFT) are used to survey potential admixing candidates that modify either the conduction-<span class="hlt">band</span> minimummore » (CBM) or valence-<span class="hlt">band</span> maximum (VBM). We consider two sets of compositions based on Lu 3B 5O 12 where B is Al, Ga, In, As, and Sb, and R 3Al 5O 12, where R is Lu, Gd, Dy, and Er. We find that admixing with various R cations does not appreciably affect the <span class="hlt">band</span> gap or <span class="hlt">band</span> edges. In contrast, substituting Al with cations of dissimilar ionic radii has a profound impact on the <span class="hlt">band</span> structure. We further show that certain dopants can be used to selectively modify only the CBM or the VBM. Specifically, Ga and In decrease the <span class="hlt">band</span> gap by lowering the CBM, while As and Sb decrease the <span class="hlt">band</span> gap by raising the VBM, the relative change in <span class="hlt">band</span> gap is quantitatively validated by HDFT. These results demonstrate a powerful approach to quickly screen the impact of dopants on the electronic structure of scintillator compounds, identifying those dopants which alter the <span class="hlt">band</span> edges in very specific ways to eliminate both electron and hole traps responsible for performance limitations. Furthermore, this approach should be broadly applicable for the optimization of electronic and optical performance for a wide range of compounds by tuning the VBM and CBM.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2008SPIE.7132E..06G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2008SPIE.7132E..06G"><span><span class="hlt">Energy</span> dependence of effective electron mass and laser-induced ionization of wide <span class="hlt">band</span>-gap solids</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gruzdev, V. E.</p> <p>2008-10-01</p> <p>Most of the traditional theoretical models of laser-induced ionization were developed under the assumption of constant effective electron mass or weak dependence of the effective mass on electron <span class="hlt">energy</span>. Those assumptions exclude from consideration all the effects resulting from significant increase of the effective mass with increasing of electron <span class="hlt">energy</span> in real the conduction <span class="hlt">band</span>. Promotion of electrons to the states with high effective mass can be done either via laserinduced electron oscillations or via electron-particle collisions. Increase of the effective mass during laser-material interactions can result in specific regimes of ionization. Performing a simple qualitative analysis by comparison of the constant-mass approximation vs realistic dependences of the effective mass on electron <span class="hlt">energy</span>, we demonstrate that the traditional ionization models provide reliable estimation of the ionization rate in a very limited domain of laser intensity and wavelength. By taking into account increase of the effective mass with electron <span class="hlt">energy</span>, we demonstrate that special regimes of high-intensity photo-ionization are possible depending on laser and material parameters. Qualitative analysis of the <span class="hlt">energy</span> dependence of the effective mass also leads to conclusion that the avalanche ionization can be stopped by the effect of electron trapping in the states with large values of the effective mass.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011APS..MARD36007W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011APS..MARD36007W"><span>Intermediate <span class="hlt">Band</span> Gap Solar Cells: The Effect of Resonant Tunneling on Delocalization</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>William, Reid; Mathew, Doty; Sanwli, Shilpa; Gammon, Dan; Bracker, Allan</p> <p>2011-03-01</p> <p>Quantum dots (QD's) have many unique properties, including tunable discrete <span class="hlt">energy</span> levels, that make them suitable for a variety of next generation photovoltaic applications. One application is an intermediate <span class="hlt">band</span> solar cell (IBSC); in which QD's are incorporated into the bulk material. The QD's are tuned to absorb low <span class="hlt">energy</span> photons that would otherwise be wasted because their <span class="hlt">energy</span> is less than the solar cell's bulk <span class="hlt">band</span> gap. Current theory concludes that identical QD's should be arranged in a superlattice to form a completely delocalized intermediate <span class="hlt">band</span> maximizing absorption of low <span class="hlt">energy</span> photons while minimizing the decrease in the efficiency of the bulk material. We use a T-matrix model to assess the feasibility of forming a delocalized <span class="hlt">band</span> given that real QD ensembles have an inhomogeneous distribution of <span class="hlt">energy</span> levels. Our results suggest that formation of a <span class="hlt">band</span> delocalized through a large QD superlattice is challenging; suggesting that the assumptions underlying present IBSC theory require reexamination. We use time-resolved photoluminescence of coupled QD's to probe the effect of delocalized states on the dynamics of absorption, <span class="hlt">energy</span> transport, and nonradiative relaxation. These results will allow us to reexamine the theoretical assumptions and determine the degree of delocalization necessary to create an efficient quantum dot-based IBSC.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018PhRvB..97s5102P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PhRvB..97s5102P"><span>Absence of a charge diffusion pole at finite <span class="hlt">energies</span> in an exactly solvable interacting flat-<span class="hlt">band</span> model in d dimensions</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Phillips, Philip W.; Setty, Chandan; Zhang, Shuyi</p> <p>2018-05-01</p> <p>Motivated by recent bounds for charge diffusion in critical matter, we investigate the following question: What sets the scale for the velocity for diffusing degrees of freedom in a scale-invariant system? To make our statements precise, we analyze the diffusion pole in an exactly solvable model for a Mott transition in the presence of a long-range interaction term. To achieve scale invariance, we limit our discussion to the flat-<span class="hlt">band</span> regime. We find in this limit that the diffusion pole, which would normally obtain at finite <span class="hlt">energy</span>, is pushed to zero <span class="hlt">energy</span>, resulting in a vanishing of the diffusion constant. This occurs even in the presence of interactions in certain limits, indicating the robustness of this result to the inclusion of a scale in the problem. Consequently, scale invariance precludes any reasonable definition of the diffusion constant. Nonetheless, we do find that a scale can be defined, albeit irrelevant to diffusion, which is the product of the squared <span class="hlt">band</span> velocity and the density of states.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018PhRvC..97d1304P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PhRvC..97d1304P"><span>Evidence of chiral <span class="hlt">bands</span> in even-even nuclei</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Petrache, C. M.; Lv, B. F.; Astier, A.; Dupont, E.; Wang, Y. K.; Zhang, S. Q.; Zhao, P. W.; Ren, Z. X.; Meng, J.; Greenlees, P. T.; Badran, H.; Cox, D. M.; Grahn, T.; Julin, R.; Juutinen, S.; Konki, J.; Pakarinen, J.; Papadakis, P.; Partanen, J.; Rahkila, P.; Sandzelius, M.; Saren, J.; Scholey, C.; Sorri, J.; Stolze, S.; Uusitalo, J.; Cederwall, B.; Aktas, Ö.; Ertoprak, A.; Liu, H.; Matta, S.; Subramaniam, P.; Guo, S.; Liu, M. L.; Zhou, X. H.; Wang, K. L.; Kuti, I.; Timár, J.; Tucholski, A.; Srebrny, J.; Andreoiu, C.</p> <p>2018-04-01</p> <p>Evidence for chiral doublet <span class="hlt">bands</span> has been observed for the first time in the even-even nucleus 136Nd. One chiral <span class="hlt">band</span> was firmly established. Four other candidates for chiral <span class="hlt">bands</span> were also identified, which can contribute to the realization of the multiple pairs of chiral doublet <span class="hlt">bands</span> (M χ D ) phenomenon. The observed <span class="hlt">bands</span> are investigated by the constrained and tilted axis cranking covariant density functional theory (TAC-CDFT). Possible configurations have been explored. The experimental <span class="hlt">energy</span> spectra, angular momenta, and B (M 1 )/B (E 2 ) values for the assigned configurations are globally reproduced by TAC-CDFT. Calculated results support the chiral interpretation of the observed <span class="hlt">bands</span>, which correspond to shapes with maximum triaxiality induced by different multiquasiparticle configurations in 136Nd.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li class="active"><span>12</span></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_12 --> <div id="page_13" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li class="active"><span>13</span></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="241"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26226296','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26226296"><span>Esaki Diodes in van der Waals Heterojunctions with Broken-Gap <span class="hlt">Energy</span> <span class="hlt">Band</span> Alignment.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Yan, Rusen; Fathipour, Sara; Han, Yimo; Song, Bo; Xiao, Shudong; Li, Mingda; Ma, Nan; Protasenko, Vladimir; Muller, David A; Jena, Debdeep; Xing, Huili Grace</p> <p>2015-09-09</p> <p>van der Waals (vdW) heterojunctions composed of two-dimensional (2D) layered materials are emerging as a solid-state materials family that exhibits novel physics phenomena that can power a range of electronic and photonic applications. Here, we present the first demonstration of an important building block in vdW solids: room temperature Esaki tunnel diodes. The Esaki diodes were realized in vdW heterostructures made of black phosphorus (BP) and tin diselenide (SnSe2), two layered semiconductors that possess a broken-gap <span class="hlt">energy</span> <span class="hlt">band</span> offset. The presence of a thin insulating barrier between BP and SnSe2 enabled the observation of a prominent negative differential resistance (NDR) region in the forward-bias current-voltage characteristics, with a peak to valley ratio of 1.8 at 300 K and 2.8 at 80 K. A weak temperature dependence of the NDR indicates electron tunneling being the dominant transport mechanism, and a theoretical model shows excellent agreement with the experimental results. Furthermore, the broken-gap <span class="hlt">band</span> alignment is confirmed by the junction photoresponse, and the phosphorus double planes in a single layer of BP are resolved in transmission electron microscopy (TEM) for the first time. Our results represent a significant advance in the fundamental understanding of vdW heterojunctions and broaden the potential applications of 2D layered materials.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26206396','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26206396"><span>Hazard <span class="hlt">banding</span> in compliance with the new Globally Harmonised System (GHS) for use in control <span class="hlt">banding</span> tools.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Arnone, Mario; Koppisch, Dorothea; Smola, Thomas; Gabriel, Stefan; Verbist, Koen; Visser, Remco</p> <p>2015-10-01</p> <p>Many control <span class="hlt">banding</span> tools use hazard <span class="hlt">banding</span> in risk assessments for the occupational handling of hazardous substances. The outcome of these assessments can be combined with advice for the required risk management measures (RMMs). The Globally Harmonised System of Classification and Labelling of Chemicals (GHS) has resulted in a change in the hazard communication elements, i.e. Hazard (H) statements instead of Risk-phrases. Hazard <span class="hlt">banding</span> schemes that depend on the old form of safety information have to be adapted to the new rules. The purpose of this publication is to outline the rationales for the assignment of hazard <span class="hlt">bands</span> to H statements under the GHS. Based on this, this publication proposes a hazard <span class="hlt">banding</span> scheme that uses the information from the safety data sheets as the basis for assignment. The assignment of hazard <span class="hlt">bands</span> tiered according to the severity of the underlying hazards supports the important principle of substitution. <span class="hlt">Additionally</span>, the set of assignment rules permits an exposure-route-specific assignment of hazard <span class="hlt">bands</span>, which is necessary for the proposed route-specific RMMs. Ideally, all control <span class="hlt">banding</span> tools should apply the same assignment rules. This GHS-compliant hazard <span class="hlt">banding</span> scheme can hopefully help to establish a unified hazard <span class="hlt">banding</span> strategy in the various control <span class="hlt">banding</span> tools. Copyright © 2015 Elsevier Inc. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012ApPhL.100f2102Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012ApPhL.100f2102Y"><span><span class="hlt">Energy</span> <span class="hlt">band</span> engineering and controlled p-type conductivity of CuAlO2 thin films by nonisovalent Cu-O alloying</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yao, Z. Q.; He, B.; Zhang, L.; Zhuang, C. Q.; Ng, T. W.; Liu, S. L.; Vogel, M.; Kumar, A.; Zhang, W. J.; Lee, C. S.; Lee, S. T.; Jiang, X.</p> <p>2012-02-01</p> <p>The electronic <span class="hlt">band</span> structure and p-type conductivity of CuAlO2 films were modified via synergistic effects of <span class="hlt">energy</span> <span class="hlt">band</span> offset and partial substitution of less-dispersive Cu+ 3d10 with Cu2+ 3d9 orbitals in the valence <span class="hlt">band</span> maximum by alloying nonisovalent Cu-O with CuAlO2 host. The Cu-O/CuAlO2 alloying films show excellent electronic properties with tunable wide direct bandgaps (˜3.46-3.87 eV); Hall measurements verify the highest hole mobilities (˜11.3-39.5 cm2/Vs) achieved thus far for CuAlO2 thin films and crystals. Top-gate thin film transistors constructed on p-CuAlO2 films were presented, and the devices showed pronounced performance with Ion/Ioff of ˜8.0 × 102 and field effect mobility of 0.97 cm2/Vs.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28265085','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28265085"><span>Understanding <span class="hlt">band</span> gaps of solids in generalized Kohn-Sham theory.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Perdew, John P; Yang, Weitao; Burke, Kieron; Yang, Zenghui; Gross, Eberhard K U; Scheffler, Matthias; Scuseria, Gustavo E; Henderson, Thomas M; Zhang, Igor Ying; Ruzsinszky, Adrienn; Peng, Haowei; Sun, Jianwei; Trushin, Egor; Görling, Andreas</p> <p>2017-03-14</p> <p>The fundamental <span class="hlt">energy</span> gap of a periodic solid distinguishes insulators from metals and characterizes low-<span class="hlt">energy</span> single-electron excitations. However, the gap in the <span class="hlt">band</span> structure of the exact multiplicative Kohn-Sham (KS) potential substantially underestimates the fundamental gap, a major limitation of KS density-functional theory. Here, we give a simple proof of a theorem: In generalized KS theory (GKS), the <span class="hlt">band</span> gap of an extended system equals the fundamental gap for the approximate functional if the GKS potential operator is continuous and the density change is delocalized when an electron or hole is added. Our theorem explains how GKS <span class="hlt">band</span> gaps from metageneralized gradient approximations (meta-GGAs) and hybrid functionals can be more realistic than those from GGAs or even from the exact KS potential. The theorem also follows from earlier work. The <span class="hlt">band</span> edges in the GKS one-electron spectrum are also related to measurable <span class="hlt">energies</span>. A linear chain of hydrogen molecules, solid aluminum arsenide, and solid argon provide numerical illustrations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5358356','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5358356"><span>Understanding <span class="hlt">band</span> gaps of solids in generalized Kohn–Sham theory</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Perdew, John P.; Yang, Weitao; Burke, Kieron; Yang, Zenghui; Gross, Eberhard K. U.; Scheffler, Matthias; Scuseria, Gustavo E.; Henderson, Thomas M.; Zhang, Igor Ying; Ruzsinszky, Adrienn; Peng, Haowei; Sun, Jianwei; Trushin, Egor; Görling, Andreas</p> <p>2017-01-01</p> <p>The fundamental <span class="hlt">energy</span> gap of a periodic solid distinguishes insulators from metals and characterizes low-<span class="hlt">energy</span> single-electron excitations. However, the gap in the <span class="hlt">band</span> structure of the exact multiplicative Kohn–Sham (KS) potential substantially underestimates the fundamental gap, a major limitation of KS density-functional theory. Here, we give a simple proof of a theorem: In generalized KS theory (GKS), the <span class="hlt">band</span> gap of an extended system equals the fundamental gap for the approximate functional if the GKS potential operator is continuous and the density change is delocalized when an electron or hole is added. Our theorem explains how GKS <span class="hlt">band</span> gaps from metageneralized gradient approximations (meta-GGAs) and hybrid functionals can be more realistic than those from GGAs or even from the exact KS potential. The theorem also follows from earlier work. The <span class="hlt">band</span> edges in the GKS one-electron spectrum are also related to measurable <span class="hlt">energies</span>. A linear chain of hydrogen molecules, solid aluminum arsenide, and solid argon provide numerical illustrations. PMID:28265085</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://eric.ed.gov/?q=kronig&id=EJ187373','ERIC'); return false;" href="https://eric.ed.gov/?q=kronig&id=EJ187373"><span>Calculation of the <span class="hlt">Energy-Band</span> Structure of the Kronig-Penney Model Using the Nearly-Free and Tightly-Bound-Electron Approximations</span></a></p> <p><a target="_blank" href="http://www.eric.ed.gov/ERICWebPortal/search/extended.jsp?_pageLabel=advanced">ERIC Educational Resources Information Center</a></p> <p>Wetsel, Grover C., Jr.</p> <p>1978-01-01</p> <p>Calculates the <span class="hlt">energy-band</span> structure of noninteracting electrons in a one-dimensional crystal using exact and approximate methods for a rectangular-well atomic potential. A comparison of the two solutions as a function of potential-well depth and ratio of lattice spacing to well width is presented. (Author/GA)</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1990JAP....67..908W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1990JAP....67..908W"><span><span class="hlt">Energy</span> <span class="hlt">band</span>-gap calculations of short-period (ZnTe)m(ZnSe)n and (ZnS)m(ZnSe)n strained-layer superlattices</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wu, Yi-hong; Fujita, Shizuo; Fujita, Shigeo</p> <p>1990-01-01</p> <p>We report on the calculations of <span class="hlt">energy</span> <span class="hlt">band</span> gaps based on the semiempirical tight-binding model for short-period (ZnTe)m(ZnSe)n and (ZnS)m(ZnSe)n strained-layer superlattices (SLSs). During the calculation, much attention has been paid to the modeling of strain effect. It is found that (ZnTe)m(ZnSe)n superlattices grown on InAs, InP, and GaAs substrates show very different electronic properties from each other, which is consistent with experimental results now available. Assuming that the emission observed for (ZnTe)m(ZnSe)n SLS originates from intrinsic luminescence, we obtain an unstrained valence-<span class="hlt">band</span> offset of 1.136±0.1 eV for this superlattice. On the other hand, the <span class="hlt">band</span> gap of (ZnS)m(ZnSe)n superlattice grown coherently on GaP is found to exhibit a much stronger structure dependence than that grown coherently on GaAs. The difference of <span class="hlt">energy</span> gap between superlattice with equal monolayers (m=n) and the corresponding alloy with equal chalcogenide composition is also discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22658055-agi-alloying-snte-boosts-thermoelectric-performance-via-simultaneous-valence-band-convergence-carrier-concentration-optimization','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22658055-agi-alloying-snte-boosts-thermoelectric-performance-via-simultaneous-valence-band-convergence-carrier-concentration-optimization"><span>AgI alloying in SnTe boosts the thermoelectric performance via simultaneous valence <span class="hlt">band</span> convergence and carrier concentration optimization</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Banik, Ananya; Biswas, Kanishka, E-mail: kanishka@jncasr.ac.in</p> <p></p> <p>SnTe, a Pb-free analogue of PbTe, was earlier assumed to be a poor thermoelectric material due to excess p-type carrier concentration and large <span class="hlt">energy</span> separation between light and heavy hole valence <span class="hlt">bands</span>. Here, we report the enhancement of the thermoelectric performance of p-type SnTe by Ag and I co-doping. AgI (1–6 mol%) alloying in SnTe modulates its electronic structure by increasing the <span class="hlt">band</span> gap of SnTe, which results in decrease in the <span class="hlt">energy</span> separation between its light and heavy hole valence <span class="hlt">bands</span>, thereby giving rise to valence <span class="hlt">band</span> convergence. <span class="hlt">Additionally</span>, iodine doping in the Te sublattice of SnTe decreases themore » excess p-type carrier concentration. Due to significant decrease in hole concentration and reduction of the <span class="hlt">energy</span> separation between light and heavy hole valence <span class="hlt">bands</span>, significant enhancement in Seebeck coefficient was achieved at the temperature range of 600–900 K for Sn{sub 1−x}Ag{sub x}Te{sub 1−x}I{sub x} samples. A maximum thermoelectric figure of merit, zT, of ~1.05 was achieved at 860 K in high quality crystalline ingot of p-type Sn{sub 0.95}Ag{sub 0.05}Te{sub 0.95}I{sub 0.05}. - Graphical abstract: Significant decrease in hole concentration and reduction of the <span class="hlt">energy</span> separation between light and heavy hole valence <span class="hlt">bands</span> resulted in a maximum thermoelectric figure of merit, zT, of ~1.05 at 860 K in high quality crystalline ingot of p-type Sn{sub 0.95}Ag{sub 0.05}Te{sub 0.95}I{sub 0.05}. - Highlights: • AgI alloying in SnTe increases the principle <span class="hlt">band</span> gap. • Hole concentration reduction and valence <span class="hlt">band</span> convergence enhances thermopower of SnTe-AgI. • A maximum zT of ~1.05 was achieved at 860 K in p-type Sn{sub 0.95}Ag{sub 0.05}Te{sub 0.95}I{sub 0.05}.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1379300-using-wannier-functions-improve-solid-band-gap-predictions-density-functional-theory','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1379300-using-wannier-functions-improve-solid-band-gap-predictions-density-functional-theory"><span>Using Wannier functions to improve solid <span class="hlt">band</span> gap predictions in density functional theory</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Ma, Jie; Wang, Lin-Wang</p> <p>2016-04-26</p> <p>Enforcing a straight-line condition of the total <span class="hlt">energy</span> upon removal/<span class="hlt">addition</span> of fractional electrons on eigen states has been successfully applied to atoms and molecules for calculating ionization potentials and electron affinities, but fails for solids due to the extended nature of the eigen orbitals. Here we have extended the straight-line condition to the removal/<span class="hlt">addition</span> of fractional electrons on Wannier functions constructed within the occupied/unoccupied subspaces. It removes the self-interaction <span class="hlt">energies</span> of those Wannier functions, and yields accurate <span class="hlt">band</span> gaps for solids compared to experiments. It does not have any adjustable parameters and the computational cost is at the DFT level.more » This method can also work for molecules, providing eigen <span class="hlt">energies</span> in good agreement with experimental ionization potentials and electron affinities. Our approach can be viewed as an alternative approach of the standard LDA+U procedure.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018TDM.....5a5008K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018TDM.....5a5008K"><span>Towards <span class="hlt">band</span> structure and <span class="hlt">band</span> offset engineering of monolayer Mo(1-x)W(x)S2 via Strain</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kim, Joon-Seok; Ahmad, Rafia; Pandey, Tribhuwan; Rai, Amritesh; Feng, Simin; Yang, Jing; Lin, Zhong; Terrones, Mauricio; Banerjee, Sanjay K.; Singh, Abhishek K.; Akinwande, Deji; Lin, Jung-Fu</p> <p>2018-01-01</p> <p>Semiconducting transition metal dichalcogenides (TMDs) demonstrate a wide range of optoelectronic properties due to their diverse elemental compositions, and are promising candidates for next-generation optoelectronics and <span class="hlt">energy</span> harvesting devices. However, effective <span class="hlt">band</span> offset engineering is required to implement practical structures with desirable functionalities. Here, we explore the pressure-induced <span class="hlt">band</span> structure evolution of monolayer WS2 and Mo0.5W0.5S2 using hydrostatic compressive strain applied in a diamond anvil cell (DAC) apparatus and theoretical calculations, in order to study the modulation of <span class="hlt">band</span> structure and explore the possibility of <span class="hlt">band</span> alignment engineering through different compositions. Higher W composition in Mo(1-x)W(x)S2 contributes to a greater pressure-sensitivity of direct <span class="hlt">band</span> gap opening, with a maximum value of 54 meV GPa-1 in WS2. Interestingly, while the conduction <span class="hlt">band</span> minima (CBMs) remains largely unchanged after the rapid gap increase, valence <span class="hlt">band</span> maxima (VBMs) significantly rise above the initial values. It is suggested that the pressure- and composition-engineering could introduce a wide variety of <span class="hlt">band</span> alignments including type I, type II, and type III heterojunctions, and allow to construct precise structures with desirable functionalities. No structural transition is observed during the pressure experiments, implying the pressure could provide selective modulation of <span class="hlt">band</span> offset.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/5070621-implementation-approximate-self-energy-correction-scheme-orthogonalized-linear-combination-atomic-orbitals-method-band-structure-calculations','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/5070621-implementation-approximate-self-energy-correction-scheme-orthogonalized-linear-combination-atomic-orbitals-method-band-structure-calculations"><span>Implementation of an approximate self-<span class="hlt">energy</span> correction scheme in the orthogonalized linear combination of atomic orbitals method of <span class="hlt">band</span>-structure calculations</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Gu, Z.; Ching, W.Y.</p> <p></p> <p>Based on the Sterne-Inkson model for the self-<span class="hlt">energy</span> correction to the single-particle <span class="hlt">energy</span> in the local-density approximation (LDA), we have implemented an approximate <span class="hlt">energy</span>-dependent and [bold k]-dependent [ital GW] correction scheme to the orthogonalized linear combination of atomic orbital-based local-density calculation for insulators. In contrast to the approach of Jenkins, Srivastava, and Inkson, we evaluate the on-site exchange integrals using the LDA Bloch functions throughout the Brillouin zone. By using a [bold k]-weighted <span class="hlt">band</span> gap [ital E][sub [ital g</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1209521-electronic-structures-bonding-configurations-band-gap-opening-properties-graphene-binding-low-concentration-fluorine','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1209521-electronic-structures-bonding-configurations-band-gap-opening-properties-graphene-binding-low-concentration-fluorine"><span>Electronic Structures, Bonding Configurations, and <span class="hlt">Band</span>-Gap-Opening Properties of Graphene Binding with Low-Concentration Fluorine</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Duan, Yuhua; Stinespring, Charter D.; Chorpening, Benjamin</p> <p>2015-06-18</p> <p>To better understand the effects of low-level fluorine in graphene-based sensors, first-principles density functional theory (DFT) with van der Waals dispersion interactions has been employed to investigate the structure and impact of fluorine defects on the electrical properties of single-layer graphene films. The results show that both graphite-2H and graphene have zero <span class="hlt">band</span> gaps. When fluorine bonds to a carbon atom, the carbon atom is pulled slightly above the graphene plane, creating what is referred to as a CF defect. The lowest-binding <span class="hlt">energy</span> state is found to correspond to two CF defects on nearest neighbor sites, with one fluorine abovemore » the carbon plane and the other below the plane. Overall this has the effect of buckling the graphene. The results further show that the <span class="hlt">addition</span> of fluorine to graphene leads to the formation of an <span class="hlt">energy</span> <span class="hlt">band</span> (BF) near the Fermi level, contributed mainly from the 2p orbitals of fluorine with a small contribution from the porbitals of the carbon. Among the 11 binding configurations studied, our results show that only in two cases does the BF serve as a conduction <span class="hlt">band</span> and open a <span class="hlt">band</span> gap of 0.37 eV and 0.24 eV respectively. The binding <span class="hlt">energy</span> decreases with decreasing fluorine concentration due to the interaction between neighboring fluorine atoms. The obtained results are useful for sensor development and nanoelectronics.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015JAP...117w4501C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015JAP...117w4501C"><span>Modeling direct <span class="hlt">band-to-band</span> tunneling: From bulk to quantum-confined semiconductor devices</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Carrillo-Nuñez, H.; Ziegler, A.; Luisier, M.; Schenk, A.</p> <p>2015-06-01</p> <p>A rigorous framework to study direct <span class="hlt">band-to-band</span> tunneling (BTBT) in homo- and hetero-junction semiconductor nanodevices is introduced. An interaction Hamiltonian coupling conduction and valence <span class="hlt">bands</span> (CVBs) is derived using a multiband envelope method. A general form of the BTBT probability is then obtained from the linear response to the "CVBs interaction" that drives the system out of equilibrium. Simple expressions in terms of the one-electron spectral function are developed to compute the BTBT current in two- and three-dimensional semiconductor structures. <span class="hlt">Additionally</span>, a two-<span class="hlt">band</span> envelope equation based on the Flietner model of imaginary dispersion is proposed for the same purpose. In order to characterize their accuracy and differences, both approaches are compared with full-<span class="hlt">band</span>, atomistic quantum transport simulations of Ge, InAs, and InAs-Si Esaki diodes. As another numerical application, the BTBT current in InAs-Si nanowire tunnel field-effect transistors is computed. It is found that both approaches agree with high accuracy. The first one is considerably easier to conceive and could be implemented straightforwardly in existing quantum transport tools based on the effective mass approximation to account for BTBT in nanodevices.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018PhyB..538..179H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PhyB..538..179H"><span><span class="hlt">Band</span> structure of an electron in a kind of periodic potentials with singularities</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hai, Kuo; Yu, Ning; Jia, Jiangping</p> <p>2018-06-01</p> <p>Noninteracting electrons in some crystals may experience periodic potentials with singularities and the governing Schrödinger equation cannot be defined at the singular points. The <span class="hlt">band</span> structure of a single electron in such a one-dimensional crystal has been calculated by using an equivalent integral form of the Schrödinger equation. Both the perturbed and exact solutions are constructed respectively for the cases of a general singular weak-periodic system and its an exactly solvable version, Kronig-Penney model. Any one of them leads to a special <span class="hlt">band</span> structure of the <span class="hlt">energy</span>-dependent parameter, which results in an effective correction to the previous <span class="hlt">energy-band</span> structure and gives a new explanation for forming the <span class="hlt">band</span> structure. The used method and obtained results could be a valuable aid in the study of <span class="hlt">energy</span> <span class="hlt">bands</span> in solid-state physics, and the new explanation may trigger investigation to different physical mechanism of electron <span class="hlt">band</span> structures.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1993PhRvB..47.3078K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1993PhRvB..47.3078K"><span><span class="hlt">Energy</span>-density and repetition-rate dependences of the KrF-excimer-laser-induced 1.9-eV emission <span class="hlt">band</span> in type-III fused silicas</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kuzuu, Nobu; Komatsu, Yoshikazu; Murahara, Masataka</p> <p>1993-02-01</p> <p>The <span class="hlt">energy</span>-density and repetition-rate dependences of the intensity of KrF-excimer-laser (5.0 eV) -induced 1.9-eV emission <span class="hlt">band</span> in type-III fused silicas synthesized under different conditions were investigated. The intensity of the 1.9-eV <span class="hlt">band</span> is proportional to the 1.7-th power of the <span class="hlt">energy</span> density and the 0.6-th power of the repetition rate of the laser pulse. The origin of these dependencies was discussed based on the trapped-oxygen-molecule model proposed by Awazu and Kawazoe; by irradiating with the excimer laser, ozone molecules are formed from dissolved oxygen molecules and 1.9-eV photons are emitted in the course of the photodecomposition of the ozone molecules. Therefore, a two-step photon-absorption process is needed to emit the 1.9-eV photon. To form the ozone molecule, diffusion of the oxygen atoms produced by the photodecomposition of the trapped oxygen molecules are needed. This model suggests that the intensity of the 1.9-eV <span class="hlt">band</span> is proportional to the square of the <span class="hlt">energy</span> density and the square root of the repetition rate; this dependency is nearly the same as that of our experimental result.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JAP...122q5102W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JAP...122q5102W"><span>Thermoelectric <span class="hlt">band</span> engineering: The role of carrier scattering</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Witkoske, Evan; Wang, Xufeng; Lundstrom, Mark; Askarpour, Vahid; Maassen, Jesse</p> <p>2017-11-01</p> <p>Complex electronic <span class="hlt">band</span> structures, with multiple valleys or <span class="hlt">bands</span> at the same or similar <span class="hlt">energies</span>, can be beneficial for thermoelectric performance, but the advantages can be offset by inter-valley and inter-<span class="hlt">band</span> scattering. In this paper, we demonstrate how first-principles <span class="hlt">band</span> structures coupled with recently developed techniques for rigorous simulation of electron-phonon scattering provide the capabilities to realistically assess the benefits and trade-offs associated with these materials. We illustrate the approach using n-type silicon as a model material and show that intervalley scattering is strong. This example shows that the convergence of valleys and <span class="hlt">bands</span> can improve thermoelectric performance, but the magnitude of the improvement depends sensitively on the relative strengths of intra- and inter-valley electron scattering. Because anisotropy of the <span class="hlt">band</span> structure also plays an important role, a measure of the benefit of <span class="hlt">band</span> anisotropy in the presence of strong intervalley scattering is presented.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22305937-probing-optical-band-gaps-nanoscale-nifeo-cofeo-epitaxial-films-high-resolution-electron-energy-loss-spectroscopy','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22305937-probing-optical-band-gaps-nanoscale-nifeo-cofeo-epitaxial-films-high-resolution-electron-energy-loss-spectroscopy"><span>Probing optical <span class="hlt">band</span> gaps at the nanoscale in NiFe₂O₄ and CoFe₂O₄ epitaxial films by high resolution electron <span class="hlt">energy</span> loss spectroscopy</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Dileep, K.; Loukya, B.; Datta, R., E-mail: ranjan@jncasr.ac.in</p> <p>2014-09-14</p> <p>Nanoscale optical <span class="hlt">band</span> gap variations in epitaxial thin films of two different spinel ferrites, i.e., NiFe₂O₄ (NFO) and CoFe₂O₄ (CFO), have been investigated by spatially resolved high resolution electron <span class="hlt">energy</span> loss spectroscopy. Experimentally, both NFO and CFO show indirect/direct <span class="hlt">band</span> gaps around 1.52 eV/2.74 and 2.3 eV, and 1.3 eV/2.31 eV, respectively, for the ideal inverse spinel configuration with considerable standard deviation in the <span class="hlt">band</span> gap values for CFO due to various levels of deviation from the ideal inverse spinel structure. Direct probing of the regions in both the systems with tetrahedral A site cation vacancy, which is distinct frommore » the ideal inverse spinel configuration, shows significantly smaller <span class="hlt">band</span> gap values. The experimental results are supported by the density functional theory based modified Becke-Johnson exchange correlation potential calculated <span class="hlt">band</span> gap values for the different cation configurations.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4845083','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4845083"><span>Data on <span class="hlt">energy-band</span>-gap characteristics of composite nanoparticles obtained by modification of the amorphous potassium polytitanate in aqueous solutions of transition metal salts</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Zimnyakov, D.A.; Sevrugin, A.V.; Yuvchenko, S.A.; Fedorov, F.S.; Tretyachenko, E.V.; Vikulova, M.A.; Kovaleva, D.S.; Krugova, E.Y.; Gorokhovsky, A.V.</p> <p>2016-01-01</p> <p>Here we present the data on the <span class="hlt">energy-band</span>-gap characteristics of composite nanoparticles produced by modification of the amorphous potassium polytitanate in aqueous solutions of different transition metal salts. <span class="hlt">Band</span> gap characteristics are investigated using diffuse reflection spectra of the obtained powders. Calculated logarithmic derivative quantity of the Kubelka–Munk function reveals a presence of local maxima in the regions 0.5–1.5 eV and 1.6–3.0 eV which correspond to <span class="hlt">band</span> gap values of the investigated materials. The values might be related to the constituents of the composite nanoparticles and intermediate products of their chemical interaction. PMID:27158654</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010PhRvB..81o3104S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010PhRvB..81o3104S"><span>Optical evidence of strong coupling between valence-<span class="hlt">band</span> holes and d -localized spins in Zn1-xMnxO</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sokolov, V. I.; Druzhinin, A. V.; Gruzdev, N. B.; Dejneka, A.; Churpita, O.; Hubicka, Z.; Jastrabik, L.; Trepakov, V.</p> <p>2010-04-01</p> <p>We report on optical-absorption study of Zn1-xMnxO (x=0-0.06) films on fused silica substrates taking special attention to the spectral range of the fundamental absorption edge (3.1-4 eV). Well-pronounced excitonic lines observed in the region 3.40-3.45 eV were found to shift to higher <span class="hlt">energies</span> with increasing Mn concentration. The optical <span class="hlt">band</span>-gap <span class="hlt">energy</span> increases with x too, reliably evidencing strong coupling between oxygen holes and localized spins of manganese ions. In the 3.1-3.3 eV region the optical-absorption curve in the manganese-contained films was found to shift to lower <span class="hlt">energies</span> with respect to that for undoped ZnO. The <span class="hlt">additional</span> absorption observed in this range is interpreted as a result of splitting of a localized Zhang-Rice-type state into the <span class="hlt">band</span> gap.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1415428-generation-highly-oblique-lower-band-chorus-via-nonlinear-three-wave-resonance','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1415428-generation-highly-oblique-lower-band-chorus-via-nonlinear-three-wave-resonance"><span>Generation of Highly Oblique Lower <span class="hlt">Band</span> Chorus Via Nonlinear Three-Wave Resonance</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Fu, Xiangrong; Gary, Stephen Peter; Reeves, Geoffrey D.; ...</p> <p>2017-09-05</p> <p>Chorus in the inner magnetosphere has been observed frequently at geomagnetically active times, typically exhibiting a two-<span class="hlt">band</span> structure with a quasi-parallel lower <span class="hlt">band</span> and an upper <span class="hlt">band</span> with a broad range of wave normal angles. But recent observations by Van Allen Probes confirm another type of lower <span class="hlt">band</span> chorus, which has a large wave normal angle close to the resonance cone angle. It has been proposed that these waves could be generated by a low-<span class="hlt">energy</span> beam-like electron component or by temperature anisotropy of keV electrons in the presence of a low-<span class="hlt">energy</span> plateau-like electron component. This paper, however, presents an alternativemore » mechanism for generation of this highly oblique lower <span class="hlt">band</span> chorus. Through a nonlinear three-wave resonance, a quasi-parallel lower <span class="hlt">band</span> chorus wave can interact with a mildly oblique upper <span class="hlt">band</span> chorus wave, producing a highly oblique quasi-electrostatic lower <span class="hlt">band</span> chorus wave. This theoretical analysis is confirmed by 2-D electromagnetic particle-in-cell simulations. Furthermore, as the newly generated waves propagate away from the equator, their wave normal angle can further increase and they are able to scatter low-<span class="hlt">energy</span> electrons to form a plateau-like structure in the parallel velocity distribution. As a result, the three-wave resonance mechanism may also explain the generation of quasi-parallel upper <span class="hlt">band</span> chorus which has also been observed in the magnetosphere.« less</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li class="active"><span>13</span></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_13 --> <div id="page_14" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li class="active"><span>14</span></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="261"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013JAP...113w3508R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013JAP...113w3508R"><span>Verification of <span class="hlt">band</span> offsets and electron effective masses in GaAsN/GaAs quantum wells: Spectroscopic experiment versus 10-<span class="hlt">band</span> k·p modeling</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ryczko, K.; Sek, G.; Sitarek, P.; Mika, A.; Misiewicz, J.; Langer, F.; Höfling, S.; Forchel, A.; Kamp, M.</p> <p>2013-06-01</p> <p>Optical transitions in GaAs1-xNx/GaAs quantum wells (QWs) have been probed by two complementary techniques, modulation spectroscopy in a form of photoreflectance and surface photovoltage spectroscopy. Transition <span class="hlt">energies</span> in QWs of various widths and N contents have been compared with the results of <span class="hlt">band</span> structure calculations based on the 10-<span class="hlt">band</span> k.p Hamiltonian. Due to the observation of higher order transitions in the measured spectra, the <span class="hlt">band</span> gap discontinuities at the GaAsN/GaAs interface and the electron effective masses could be determined, both treated as semi-free parameters to get the best matching between the theoretical and experimental <span class="hlt">energies</span>. We have obtained the chemical conduction <span class="hlt">band</span> offset values of 86% for x = 1.2% and 83% for x = 2.2%, respectively. For these determined <span class="hlt">band</span> offsets, the electron effective masses equal to about 0.09 mo in QWs with 1.2% N and 0.15 mo for the case of larger N content of 2.2%.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20010092183&hterms=al+gore&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dal%2Bgore','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20010092183&hterms=al+gore&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dal%2Bgore"><span>Improvement of the Database on the 1.13-microns <span class="hlt">Band</span> of Water Vapor</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Giver, Lawrence P.; Schwenke, David W.; Chackerian, Charles, Jr.; Varanasi, Prasad; Freedman, Richard S.; Gore, Warren J. (Technical Monitor)</p> <p>2000-01-01</p> <p>Corrections have recently been reported (Giver et al.) on the short-wave (visible and near-infrared) line intensities of water vapor that were catalogued in the spectroscopic database known as HITRAN. These updates have been posted on www.hitran.com, and are being used to reanalyze the polar stratospheric absorption in the 0.94 microns <span class="hlt">band</span> as observed in POAM. We are currently investigating <span class="hlt">additional</span> improvement in the 1.13 microns <span class="hlt">band</span> using data obtained by us with an absorption path length of 1.107 km and 4 torr of water vapor and the ab initio line list of Partridge and Schwenke (needs ref). We are proposing the following four types of improvement of the HITRAN database in this region: 1) HITRAN has nearly 200 lines in this region without proper assignments of rotational quantum levels. Nearly all of them can now be assigned. 2) We have measured positions of the observable H2O-17 and H2O-18 lines. These lines in HITRAN currently have approximate positions based upon rather aged computations. 3) Some <span class="hlt">additional</span> lines are observed and assigned which should be included in the database. 4) Corrections are necessary for the lower state <span class="hlt">energies</span> E" for the HITRAN lines of the 121-010 "hot" <span class="hlt">band</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1399113-methodology-wide-band-gap-device-dynamic-characterization','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1399113-methodology-wide-band-gap-device-dynamic-characterization"><span>Methodology for Wide <span class="hlt">Band</span>-Gap Device Dynamic Characterization</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Zhang, Zheyu; Guo, Ben; Wang, Fei Fred; ...</p> <p>2017-01-19</p> <p>Here, the double pulse test (DPT) is a widely accepted method to evaluate the dynamic behavior of power devices. Considering the high switching-speed capability of wide <span class="hlt">band</span>-gap devices, the test results are very sensitive to the alignment of voltage and current (V-I) measurements. Also, because of the shoot-through current induced by Cdv/dt (i.e., cross-talk), the switching losses of the nonoperating switch device in a phase-leg must be considered in <span class="hlt">addition</span> to the operating device. This paper summarizes the key issues of the DPT, including components and layout design, measurement considerations, grounding effects, and data processing. <span class="hlt">Additionally</span>, a practical method ismore » proposed for phase-leg switching loss evaluation by calculating the difference between the input <span class="hlt">energy</span> supplied by a dc capacitor and the output <span class="hlt">energy</span> stored in a load inductor. Based on a phase-leg power module built with 1200-V/50-A SiC MOSFETs, the test results show that this method can accurately evaluate the switching loss of both the upper and lower switches by detecting only one switching current and voltage, and it is immune to V-I timing misalignment errors.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018PhRvB..97p5130E','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PhRvB..97p5130E"><span>Location of the valence <span class="hlt">band</span> maximum in the <span class="hlt">band</span> structure of anisotropic 1 T'-ReSe2</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Eickholt, P.; Noky, J.; Schwier, E. F.; Shimada, K.; Miyamoto, K.; Okuda, T.; Datzer, C.; Drüppel, M.; Krüger, P.; Rohlfing, M.; Donath, M.</p> <p>2018-04-01</p> <p>Transition-metal dichalcogenides (TMDCs) are a focus of current research due to their fascinating optical and electronic properties with possible technical applications. ReSe2 is an interesting material of the TMDC family, with unique anisotropic properties originating from its distorted 1 T structure (1 T '). To develop a fundamental understanding of the optical and electric properties, we studied the underlying electronic structure with angle-resolved photoemission (ARPES) as well as <span class="hlt">band</span>-structure calculations within the density functional theory (DFT)-local density approximation (LDA) and GdW approximations. We identified the Γ ¯M¯1 direction, which is perpendicular to the a axis, as a distinct direction in k space with the smallest bandwidth of the highest valence <span class="hlt">band</span>. Using photon-<span class="hlt">energy</span>-dependent ARPES, two valence <span class="hlt">band</span> maxima are identified within experimental limits of about 50 meV: one at the high-symmetry point Z , and a second one at a non-high-symmetry point in the Brillouin zone. Thus, the position in k space of the global valence <span class="hlt">band</span> maximum is undecided experimentally. Theoretically, an indirect <span class="hlt">band</span> gap is predicted on a DFT-LDA level, while quasiparticle corrections lead to a direct <span class="hlt">band</span> gap at the Z point.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018PhRvB..97d1203Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PhRvB..97d1203Z"><span>Flat <span class="hlt">band</span> in disorder-driven non-Hermitian Weyl semimetals</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zyuzin, A. A.; Zyuzin, A. Yu.</p> <p>2018-01-01</p> <p>We study the interplay of disorder and <span class="hlt">band</span>-structure topology in a Weyl semimetal with a tilted conical spectrum around the Weyl points. The spectrum of particles is given by the eigenvalues of a non-Hermitian matrix, which contains contributions from a Weyl Hamiltonian and complex self-<span class="hlt">energy</span> due to electron elastic scattering on disorder. We find that the tilt-induced matrix structure of the self-<span class="hlt">energy</span> gives rise to either a flat <span class="hlt">band</span> or a nodal line segment at the interface of the electron and hole pockets in the bulk <span class="hlt">band</span> structure of type-II Weyl semimetals depending on the Weyl cone inclination. For the tilt in a single direction in momentum space, each Weyl point expands into a flat <span class="hlt">band</span> lying on the plane, which is transverse to the direction of the tilt. The spectrum of the flat <span class="hlt">band</span> is fully imaginary and is separated from the in-plane dispersive part of the spectrum by the "exceptional nodal ring" where the matrix of the Green's function in momentum-frequency space is defective. The tilt in two directions might shrink a flat <span class="hlt">band</span> into a nodal line segment with "exceptional edge points." We discuss the connection to the non-Hermitian topological theory.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1988PhRvB..38.7723L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1988PhRvB..38.7723L"><span>Transition-metal impurities in semiconductors and heterojunction <span class="hlt">band</span> lineups</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Langer, Jerzy M.; Delerue, C.; Lannoo, M.; Heinrich, Helmut</p> <p>1988-10-01</p> <p>The validity of a recent proposal that transition-metal impurity levels in semiconductors may serve as a reference in <span class="hlt">band</span> alignment in semiconductor heterojunctions is positively verified by using the most recent data on <span class="hlt">band</span> offsets in the following lattice-matched heterojunctions: Ga1-xAlxAs/GaAs, In1-xGaxAsyP1-y/InP, In1-xGaxP/GaAs, and Cd1-xHgxTe/CdTe. The alignment procedure is justified theoretically by showing that transition-metal <span class="hlt">energy</span> levels are effectively pinned to the average dangling-bond <span class="hlt">energy</span> level, which serves as the reference level for the heterojunction <span class="hlt">band</span> alignment. Experimental and theoretical arguments showing that an increasingly popular notion on transition-metal <span class="hlt">energy</span>-level pinning to the vacuum level is unjustified and must be abandoned in favor of the internal-reference rule proposed recently [J. M. Langer and H. Heinrich, Phys. Rev. Lett. 55, 1414 (1985)] are presented.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015RuPhJ..58..179G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015RuPhJ..58..179G"><span>Influence of Water Vapors and Hydrogen on the <span class="hlt">Energy</span> <span class="hlt">Band</span> Bending in the SnO2 Microcrystals of Polycrystalline Tin Dioxide Films</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gaman, V. I.; Almaev, A. V.; Sevast'yanov, E. Yu.; Maksimova, N. K.</p> <p>2015-06-01</p> <p>The results of studying the dependence of the <span class="hlt">energy</span> <span class="hlt">band</span> bending at the interface of contacting SnO2 microcrystals in the polycrystalline tin dioxide film on the humidity level of clean air and hydrogen concentration in the gas mixture of clean air + H2 are presented. The experimental results showed that the bending of <span class="hlt">energy</span> <span class="hlt">bands</span> in SnO2 is decreased under exposure to the water vapors and molecular hydrogen. The presence of two types of the adsorption centers for water molecules on the surface of SnO2 is found. It is shown that at the absolute humidity of the gas mixture above 12 g/m3, the H2O and H2 molecules are adsorbed on the same centers, whose surface density is of 1012 сm-2 at a concentration of donor impurity in SnO2 equal to 1018 сm-3.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/21455104-near-infrared-thermal-emission-from-tres-ks-band-detection-band-upper-limit-depth-secondary-eclipse','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/21455104-near-infrared-thermal-emission-from-tres-ks-band-detection-band-upper-limit-depth-secondary-eclipse"><span>NEAR-INFRARED THERMAL EMISSION FROM TrES-3b: A Ks-<span class="hlt">BAND</span> DETECTION AND AN H-<span class="hlt">BAND</span> UPPER LIMIT ON THE DEPTH OF THE SECONDARY ECLIPSE</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Croll, Bryce; Jayawardhana, Ray; Fortney, Jonathan J.</p> <p>2010-08-01</p> <p>We present H- and Ks-<span class="hlt">band</span> photometry bracketing the secondary eclipse of the hot Jupiter TrES-3b using the Wide-field Infrared Camera on the Canada-France-Hawaii Telescope. We detect the secondary eclipse of TrES-3b with a depth of 0.133{sup +0.018}{sub -0.016}% in the Ks <span class="hlt">band</span> (8{sigma})-a result that is in sharp contrast to the eclipse depth reported by de Mooij and Snellen. We do not detect its thermal emission in the H <span class="hlt">band</span>, but place a 3{sigma} limit of 0.051% on the depth of the secondary eclipse in this <span class="hlt">band</span>. A secondary eclipse of this depth in Ks requires very efficient day-to-nightside redistributionmore » of heat and nearly isotropic reradiation, a conclusion that is in agreement with longer wavelength, mid-infrared Spitzer observations. Our 3{sigma} upper limit on the depth of our H-<span class="hlt">band</span> secondary eclipse also argues for very efficient redistribution of heat and suggests that the atmospheric layer probed by these observations may be well homogenized. However, our H-<span class="hlt">band</span> upper limit is so constraining that it suggests the possibility of a temperature inversion at depth, or an absorbing molecule, such as methane, that further depresses the emitted flux at this wavelength. The combination of our near-infrared measurements and those obtained with Spitzer suggests that TrES-3b displays a near-isothermal dayside atmospheric temperature structure, whose spectrum is well approximated by a blackbody. We emphasize that our strict H-<span class="hlt">band</span> limit is in stark disagreement with the best-fit atmospheric model that results from longer wavelength observations only, thus highlighting the importance of near-infrared observations at multiple wavelengths, in <span class="hlt">addition</span> to those returned by Spitzer in the mid-infrared, to facilitate a comprehensive understanding of the <span class="hlt">energy</span> budgets of transiting exoplanets.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28205063','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28205063"><span>Excitation <span class="hlt">energy</span> transfer from the bacteriochlorophyll Soret <span class="hlt">band</span> to carotenoids in the LH2 light-harvesting complex from Ectothiorhodospira haloalkaliphila is negligible.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Razjivin, A P; Lukashev, E P; Kompanets, V O; Kozlovsky, V S; Ashikhmin, A A; Chekalin, S V; Moskalenko, A A; Paschenko, V Z</p> <p>2017-09-01</p> <p>Pathways of intramolecular conversion and intermolecular electronic excitation <span class="hlt">energy</span> transfer (EET) in the photosynthetic apparatus of purple bacteria remain subject to debate. Here we experimentally tested the possibility of EET from the bacteriochlorophyll (BChl) Soret <span class="hlt">band</span> to the singlet S 2 level of carotenoids using femtosecond pump-probe measurements and steady-state fluorescence excitation and absorption measurements in the near-ultraviolet and visible spectral ranges. The efficiency of EET from the Soret <span class="hlt">band</span> of BChl to S 2 of the carotenoids in light-harvesting complex LH2 from the purple bacterium Ectothiorhodospira haloalkaliphila appeared not to exceed a few percent.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014SSCom.190...44G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014SSCom.190...44G"><span>Excitonic and <span class="hlt">band-band</span> transitions of Cu2ZnSiS4 determined from reflectivity spectra</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Guc, M.; Levcenko, S.; Dermenji, L.; Gurieva, G.; Schorr, S.; Syrbu, N. N.; Arushanov, E.</p> <p>2014-07-01</p> <p>Exciton spectra of Cu2ZnSiS4 single crystals are investigated by reflection spectroscopy at 10 and 300 K for light polarized perpendicular (E⊥c) and parallel (E∥c) to the optical axis. The parameters of the excitons and dielectric constant are determined. The free carriers effective masses have been estimated. The room temperature reflectivity spectra at photon <span class="hlt">energies</span> higher than the fundamental <span class="hlt">band</span> gap in the polarization Е⊥с and E∥с were measured and related to the electronic <span class="hlt">band</span> structure of Cu2ZnSiS4.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2009NuPhA.825...16B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009NuPhA.825...16B"><span><span class="hlt">Band</span> structures in near spherical 138Ce</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bhattacharjee, T.; Chanda, S.; Bhattacharyya, S.; Basu, S. K.; Bhowmik, R. K.; Das, J. J.; Pramanik, U. Datta; Ghugre, S. S.; Madhavan, N.; Mukherjee, A.; Mukherjee, G.; Muralithar, S.; Singh, R. P.</p> <p>2009-06-01</p> <p>The high spin states of N=80138Ce have been populated in the fusion evaporation reaction 130Te( 12C, 4n) 138Ce at E=65 MeV. The γ transitions belonging to various <span class="hlt">band</span> structures were detected and characterized using an array of five Clover Germanium detectors. The level scheme has been established up to a maximum spin and excitation <span class="hlt">energy</span> of 23 ℏ and 9511.3 keV, respectively, by including 53 new transitions. The negative parity ΔI=1 <span class="hlt">band</span>, developed on the 6536.3 keV 15 level, has been conjectured to be a magnetic rotation <span class="hlt">band</span> following a semiclassical analysis and comparing the systematics of similar <span class="hlt">bands</span> in the neighboring nuclei. The said <span class="hlt">band</span> is proposed to have a four quasiparticle configuration of [πgh]⊗[. Other <span class="hlt">band</span> structures are interpreted in terms of multi-quasiparticle configurations, based on Total Routhian Surface (TRS) calculations. For the low and medium spin states, a shell model calculation using a realistic two body interaction has been performed using the code OXBASH.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22596910-layer-specific-optical-band-gap-measurement-nanoscale-mos-sub-res-sub-van-der-waals-compounds-high-resolution-electron-energy-loss-spectroscopy','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22596910-layer-specific-optical-band-gap-measurement-nanoscale-mos-sub-res-sub-van-der-waals-compounds-high-resolution-electron-energy-loss-spectroscopy"><span>Layer specific optical <span class="hlt">band</span> gap measurement at nanoscale in MoS{sub 2} and ReS{sub 2} van der Waals compounds by high resolution electron <span class="hlt">energy</span> loss spectroscopy</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Dileep, K., E-mail: dileep@jncasr.ac.in, E-mail: ranjan@jncasr.ac.in; Sahu, R.; Datta, R., E-mail: dileep@jncasr.ac.in, E-mail: ranjan@jncasr.ac.in</p> <p>2016-03-21</p> <p>Layer specific direct measurement of optical <span class="hlt">band</span> gaps of two important van der Waals compounds, MoS{sub 2} and ReS{sub 2}, is performed at nanoscale by high resolution electron <span class="hlt">energy</span> loss spectroscopy. For monolayer MoS{sub 2}, the twin excitons (1.8 and 1.95 eV) originating at the K point of the Brillouin zone are observed. An indirect <span class="hlt">band</span> gap of 1.27 eV is obtained from the multilayer regions. Indirect to direct <span class="hlt">band</span> gap crossover is observed which is consistent with the previously reported strong photoluminescence from the monolayer MoS{sub 2}. For ReS{sub 2}, the <span class="hlt">band</span> gap is direct, and a value of 1.52 andmore » 1.42 eV is obtained for the monolayer and multilayer, respectively. The <span class="hlt">energy</span> loss function is dominated by features due to high density of states at both the valence and conduction <span class="hlt">band</span> edges, and the difference in analyzing <span class="hlt">band</span> gap with respect to ZnO is highlighted. Crystalline 1T ReS{sub 2} forms two dimensional chains like superstructure due to the clustering between four Re atoms. The results demonstrate the power of HREELS technique as a nanoscale optical absorption spectroscopy tool.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20040081070&hterms=use+remote+sensing&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Duse%2Bremote%2Bsensing','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20040081070&hterms=use+remote+sensing&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Duse%2Bremote%2Bsensing"><span>Use of IRI to Model the Effect of Ionosphere Emission on Earth Remote Sensing at L-<span class="hlt">Band</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Abraham, Saji; LeVine, David M.</p> <p>2004-01-01</p> <p>Microwave remote sensing in the window at 1.413 GHz (L-<span class="hlt">band</span>) set aside for passive use only is important for monitoring sea surface salinity and soil moisture. These parameters are important for understanding ocean dynamics and <span class="hlt">energy</span> exchange between the surface and atmosphere, and both NASA and ESA plan to launch satellite sensors to monitor these parameters at L-<span class="hlt">band</span> (Aquarius, Hydros and SMOS). The ionosphere is an important source of error for passive remote sensing at this frequency. In <span class="hlt">addition</span> to Faraday rotation, emission from the ionosphere is also a potential source of error at L-<span class="hlt">band</span>. As an aid for correcting for emission, a regression model is presented that relates ionosphere emission to the integrated electron density (TEC). The goal is to use TEC from sources such as TOPEX, JASON or GPS to obtain estimates of emission over the oceans where the electron density profiles needed to compute emission are not available. In <span class="hlt">addition</span>, data will also be presented to evaluate the use of the IRI for computing emission over the ocean.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014JKPS...64..205L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014JKPS...64..205L"><span>Design study of an S-<span class="hlt">band</span> RF cavity of a dual-<span class="hlt">energy</span> electron LINAC for the CIS</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lee, Byeong-No; Park, Hyungdal; Song, Ki-baek; Li, Yonggui; Lee, Byung Cheol; Cha, Sung-su; Lee, Jong-Chul; Shin, Seung-Wook; Chai, Jong-seo</p> <p>2014-01-01</p> <p>The design of a resonance frequency (RF) cavity for the dual-<span class="hlt">energy</span> S-<span class="hlt">band</span> electron linear accelerator (LINAC) has been carried out for the cargo inspection system (CIS). This Standing-wave-type RF cavity is operated at a frequency under the 2856-MHz resonance frequency and generates electron beams of 9 MeV (high mode) and 6 MeV (low mode). The electrons are accelerated from the initial <span class="hlt">energy</span> of the electron gun to the target <span class="hlt">energy</span> (9 or 6 MeV) inside the RF cavity by using the RF power transmitted from a 5.5-MW-class klystron. Then, electron beams with a 1-kW average power (both high mode and low mode) bombard an X-ray target a 2-mm spot size. The proposed accelerating gradient was 13 MV/m, and the designed Q value was about 7100. On going research on 15-MeV non-destructive inspections for military or other applications is presented.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016ChPhC..40b4102Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016ChPhC..40b4102Z"><span>Chiral geometry in multiple chiral doublet <span class="hlt">bands</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zhang, Hao; Chen, Qibo</p> <p>2016-02-01</p> <p>The chiral geometry of multiple chiral doublet <span class="hlt">bands</span> with identical configuration is discussed for different triaxial deformation parameters γ in the particle rotor model with . The <span class="hlt">energy</span> spectra, electromagnetic transition probabilities B(M1) and B(E2), angular momenta, and K-distributions are studied. It is demonstrated that the chirality still remains not only in the yrast and yrare <span class="hlt">bands</span>, but also in the two higher excited <span class="hlt">bands</span> when γ deviates from 30°. The chiral geometry relies significantly on γ, and the chiral geometry of the two higher excited partner <span class="hlt">bands</span> is not as good as that of the yrast and yrare doublet <span class="hlt">bands</span>. Supported by Plan Project of Beijing College Students’ Scientific Research and Entrepreneurial Action, Major State 973 Program of China (2013CB834400), National Natural Science Foundation of China (11175002, 11335002, 11375015, 11461141002), National Fund for Fostering Talents of Basic Science (NFFTBS) (J1103206), Research Fund for Doctoral Program of Higher Education (20110001110087) and China Postdoctoral Science Foundation (2015M580007)</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/1347098-multi-band-uncertainty-set-based-robust-scuc-spatial-temporal-budget-constraints','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/1347098-multi-band-uncertainty-set-based-robust-scuc-spatial-temporal-budget-constraints"><span>A Multi-<span class="hlt">Band</span> Uncertainty Set Based Robust SCUC With Spatial and Temporal Budget Constraints</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Dai, Chenxi; Wu, Lei; Wu, Hongyu</p> <p>2016-11-01</p> <p>The dramatic increase of renewable <span class="hlt">energy</span> resources in recent years, together with the long-existing load forecast errors and increasingly involved price sensitive demands, has introduced significant uncertainties into power systems operation. In order to guarantee the operational security of power systems with such uncertainties, robust optimization has been extensively studied in security-constrained unit commitment (SCUC) problems, for immunizing the system against worst uncertainty realizations. However, traditional robust SCUC models with single-<span class="hlt">band</span> uncertainty sets may yield over-conservative solutions in most cases. This paper proposes a multi-<span class="hlt">band</span> robust model to accurately formulate various uncertainties with higher resolution. By properly tuning <span class="hlt">band</span> intervalsmore » and weight coefficients of individual <span class="hlt">bands</span>, the proposed multi-<span class="hlt">band</span> robust model can rigorously and realistically reflect spatial/temporal relationships and asymmetric characteristics of various uncertainties, and in turn could effectively leverage the tradeoff between robustness and economics of robust SCUC solutions. The proposed multi-<span class="hlt">band</span> robust SCUC model is solved by Benders decomposition (BD) and outer approximation (OA), while taking the advantage of integral property of the proposed multi-<span class="hlt">band</span> uncertainty set. In <span class="hlt">addition</span>, several accelerating techniques are developed for enhancing the computational performance and the convergence speed. Numerical studies on a 6-bus system and the modified IEEE 118-bus system verify the effectiveness of the proposed robust SCUC approach for enhancing uncertainty modeling capabilities and mitigating conservativeness of the robust SCUC solution.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19930068347&hterms=2060&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3D2060','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19930068347&hterms=2060&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3D2060"><span>Infrared radiation parameterizations for the minor CO2 <span class="hlt">bands</span> and for several CFC <span class="hlt">bands</span> in the window region</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Kratz, David P.; Chou, Ming-Dah; Yan, Michael M.-H.</p> <p>1993-01-01</p> <p>Fast and accurate parameterizations have been developed for the transmission functions of the CO2 9.4- and 10.4-micron <span class="hlt">bands</span>, as well as the CFC-11, CFC-12, and CFC-22 <span class="hlt">bands</span> located in the 8-12-micron region. The parameterizations are based on line-by-line calculations of transmission functions for the CO2 <span class="hlt">bands</span> and on high spectral resolution laboratory measurements of the absorption coefficients for the CFC <span class="hlt">bands</span>. Also developed are the parameterizations for the H2O transmission functions for the corresponding spectral <span class="hlt">bands</span>. Compared to the high-resolution calculations, fluxes at the tropopause computed with the parameterizations are accurate to within 10 percent when overlapping of gas absorptions within a <span class="hlt">band</span> is taken into account. For individual gas absorption, the accuracy is of order 0-2 percent. The climatic effects of these trace gases have been studied using a zonally averaged multilayer <span class="hlt">energy</span> balance model, which includes seasonal cycles and a simplified deep ocean. With the trace gas abundances taken to follow the Intergovernmental Panel on Climate Change Low Emissions 'B' scenario, the transient response of the surface temperature is simulated for the period 1900-2060.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1258569','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1258569"><span>Wild <span class="hlt">Band</span> Edges: The Role of Bandgap Grading and <span class="hlt">Band</span>-Edge Fluctuations in High-Efficiency Chalcogenide Devices: Preprint</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Repins, Ingrid; Mansfield, Lorelle; Kanevce, Ana</p> <p></p> <p><span class="hlt">Band</span>-edge effects -- including grading, electrostatic fluctuations, bandgap fluctuations, and <span class="hlt">band</span> tails -- affect chalcogenide device efficiency. These effects now require more careful consideration as efficiencies increase beyond 20%. Several aspects of the relationships between <span class="hlt">band</span>-edge phenomena and device performance for NREL absorbers are examined. For Cu(In,Ga)Se2 devices, recent increases in diffusion length imply changes to optimum bandgap profile. The origin, impact, and modification of electrostatic and bandgap fluctuations are also discussed. The application of the same principles to devices based on CdTe, kesterites, and emerging absorbers (Cu2SnS3, CuSbS2), considering differences in materials properties and defect formation <span class="hlt">energies</span>, is examined.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018AAS...23133002P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AAS...23133002P"><span>Studying Notable Debris Disks In L-<span class="hlt">band</span> with the Vortex Coronagraph</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Patel, Rahul; Beichman, Charles; Choquet, Elodie; Mawet, Dimitri; Meshkat, Tiffany; ygouf, marie</p> <p>2018-01-01</p> <p>Resolved images of circumstellar disks are integral to our understanding of planetary systems, as the micron sized dust grains that comprise the disk are born from the collisional grinding of planetesimals by larger planets in the system. Resolved images are essential to determining grain properties that might otherwise be degenerate from analyzing the star’s spectral <span class="hlt">energy</span> distribution. Though the majority of scattered light images of disks are obtained at optical and near-IR wavelengths, only a few have been imaged in the thermal IR at L-<span class="hlt">band</span>. Probing the spatial features of disks at L-<span class="hlt">band</span> opens up the possibility of constraining <span class="hlt">additional</span> grain properties, such as water/ice features.Here, we present the results of our effort to image the disks of a few notable systems at L-<span class="hlt">band</span> using the NIRC2 imager at Keck, in conjunction with the newly commissioned vector vortex coronagraph. The vortex, along with the QACITS fine guiding program installed at Keck, enables us to probe the small ~lambda/D angular separations of these systems, and reach contrasts of 1/100,000. We will discuss the systems that have been imaged, and lessons learned while imaging in L-<span class="hlt">band</span>. Our analysis of these disks reveal features previously unseen, and will lay the foundation for followup studies by missions such as JWST at similar wavelengths from space.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011ApPhL..98y3502H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011ApPhL..98y3502H"><span><span class="hlt">Band</span> alignment at the Cu2ZnSn(SxSe1-x)4/CdS interface</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Haight, Richard; Barkhouse, Aaron; Gunawan, Oki; Shin, Byungha; Copel, Matt; Hopstaken, Marinus; Mitzi, David B.</p> <p>2011-06-01</p> <p><span class="hlt">Energy</span> <span class="hlt">band</span> alignments between CdS and Cu2ZnSn(SxSe1-x)4 (CZTSSe) grown via solution-based and vacuum-based deposition routes were studied as a function of the [S]/[S+Se] ratio with femtosecond laser ultraviolet photoelectron spectroscopy, photoluminescence, medium <span class="hlt">energy</span> ion scattering, and secondary ion mass spectrometry. <span class="hlt">Band</span> bending in the underlying CZTSSe layer was measured via pump/probe photovoltage shifts of the photoelectron spectra and offsets were determined with photoemission under flat <span class="hlt">band</span> conditions. Increasing the S content of the CZTSSe films produces a valence edge shift to higher binding <span class="hlt">energy</span> and increases the CZTSSe <span class="hlt">band</span> gap. In all cases, the CdS conduction <span class="hlt">band</span> offsets were spikes.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li class="active"><span>14</span></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_14 --> <div id="page_15" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li class="active"><span>15</span></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="281"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016MRE.....3i5005Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016MRE.....3i5005Y"><span>The dependence of the tunneling characteristic on the electronic <span class="hlt">energy</span> <span class="hlt">bands</span> and the carrier’s states of Graphene superlattice</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yang, C. H.; Shen, G. Z.; Ao, Z. M.; Xu, Y. W.</p> <p>2016-09-01</p> <p>Using the transfer matrix method, the carrier tunneling properties in graphene superlattice generated by the Thue-Morse sequence and Kolakoski sequence are investigated. The positions and strength of the transmission can be modulated by the barrier structures, the incident <span class="hlt">energy</span> and angle, the height and width of the potential. These carriers tunneling characteristic can be understood from the <span class="hlt">energy</span> <span class="hlt">band</span> structures in the corresponding superlattice systems and the carrier’s states in well/barriers. The transmission peaks above the critical incident angle rely on the carrier’s resonance in the well regions. The structural diversity can modulate the electronic and transport properties, thus expanding its applications.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/15218254','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/15218254"><span>The hierarchically organized splitting of chromosome <span class="hlt">bands</span> into sub-<span class="hlt">bands</span> analyzed by multicolor <span class="hlt">banding</span> (MCB).</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Lehrer, H; Weise, A; Michel, S; Starke, H; Mrasek, K; Heller, A; Kuechler, A; Claussen, U; Liehr, T</p> <p>2004-01-01</p> <p>To clarify the nature of chromosome sub-<span class="hlt">bands</span> in more detail, the multicolor <span class="hlt">banding</span> (MCB) probe-set for chromosome 5 was hybridized to normal metaphase spreads of GTG <span class="hlt">band</span> levels at approximately 850, approximately 550, approximately 400 and approximately 300. It could be observed that as the chromosomes became shorter, more of the initial 39 MCB pseudo-colors disappeared, ending with 18 MCB pseudo-colored <span class="hlt">bands</span> at the approximately 300-<span class="hlt">band</span> level. The hierarchically organized splitting of <span class="hlt">bands</span> into sub-<span class="hlt">bands</span> was analyzed by comparing the disappearance or appearance of pseudo-color <span class="hlt">bands</span> of the four different <span class="hlt">band</span> levels. The regions to split first are telomere-near, centromere-near and in 5q23-->q31, followed by 5p15, 5p14, and all GTG dark <span class="hlt">bands</span> in 5q apart from 5q12 and 5q32 and finalized by sub-<span class="hlt">band</span> building in 5p15.2, 5q21.2-->q21.3, 5q23.1 and 5q34. The direction of <span class="hlt">band</span> splitting towards the centromere or the telomere could be assigned to each <span class="hlt">band</span> separately. Pseudo-colors assigned to GTG-light <span class="hlt">bands</span> were resistant to <span class="hlt">band</span> splitting. These observations are in concordance with the recently proposed concept of chromosome region-specific protein swelling. Copyright 2003 S. Karger AG, Basel</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2007APS..MARN38005J','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2007APS..MARN38005J"><span>Sinc or Sine? The <span class="hlt">Band</span> Excitation Method and <span class="hlt">Energy</span> Dissipation Measurements by SPM</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Jesse, Stephen; Kalinin, Sergei</p> <p>2007-03-01</p> <p>Quantitative <span class="hlt">energy</span> dissipation measurements in force-based SPM is the key to understanding fundamental mechanisms of <span class="hlt">energy</span> transformations on the nanoscale, molecular, and atomic levels. To date, these measurements are invariably based on either phase and amplitude detection in constant frequency mode, or as amplitude detection in frequency-tracking mode. The analysis in both cases implicitly assumes that amplitude is inversely proportional to the Q-factor and is not applicable when the driving force is position dependent, as is the case for virtually all SPM measurements. All current SPM methods sample only a single frequency in the Fourier domain of the system. Thus, only two out of three parameters (amplitude, resonance, and Q) can be determined independently. Here, we developed and implemented a new approach for SPM detection based on the excitation and detection of a signal having a finite amplitude over a selected region in the Fourier domain and allows simultaneous determination of all three parameters. This <span class="hlt">band</span> excitation method allows acquisition of the local spectral response at a 10ms/pixel rate, compatible with fast imaging, and is illustrated for electromechanical and mechanical imaging and force-distance spectroscopy. The BE method thus represents a new paradigm in SPM, beyond traditional single-frequency excitation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27364116','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27364116"><span><span class="hlt">Band</span>-like transport in highly crystalline graphene films from defective graphene oxides.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Negishi, R; Akabori, M; Ito, T; Watanabe, Y; Kobayashi, Y</p> <p>2016-07-01</p> <p>The electrical transport property of the reduced graphene oxide (rGO) thin-films synthesized from defective GO through thermal treatment in a reactive ethanol environment at high temperature above 1000 °C shows a <span class="hlt">band</span>-like transport with small thermal activation <span class="hlt">energy</span> (Ea~10 meV) that occurs during high carrier mobility (~210 cm(2)/Vs). Electrical and structural analysis using X-ray absorption fine structure, the valence <span class="hlt">band</span> photo-electron, Raman spectra and transmission electron microscopy indicate that a high temperature process above 1000 °C in the ethanol environment leads to an extraordinary expansion of the conjugated π-electron system in rGO due to the efficient restoration of the graphitic structure. We reveal that Ea decreases with the increasing density of states near the Fermi level due to the expansion of the conjugated π-electron system in the rGO. This means that Ea corresponds to the <span class="hlt">energy</span> gap between the top of the valence <span class="hlt">band</span> and the bottom of the conduction <span class="hlt">band</span>. The origin of the <span class="hlt">band</span>-like transport can be explained by the carriers, which are more easily excited into the conduction <span class="hlt">band</span> due to the decreasing <span class="hlt">energy</span> gap with the expansion of the conjugated π-electron system in the rGO.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016NatSR...628936N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016NatSR...628936N"><span><span class="hlt">Band</span>-like transport in highly crystalline graphene films from defective graphene oxides</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Negishi, R.; Akabori, M.; Ito, T.; Watanabe, Y.; Kobayashi, Y.</p> <p>2016-07-01</p> <p>The electrical transport property of the reduced graphene oxide (rGO) thin-films synthesized from defective GO through thermal treatment in a reactive ethanol environment at high temperature above 1000 °C shows a <span class="hlt">band</span>-like transport with small thermal activation <span class="hlt">energy</span> (Ea~10 meV) that occurs during high carrier mobility (~210 cm2/Vs). Electrical and structural analysis using X-ray absorption fine structure, the valence <span class="hlt">band</span> photo-electron, Raman spectra and transmission electron microscopy indicate that a high temperature process above 1000 °C in the ethanol environment leads to an extraordinary expansion of the conjugated π-electron system in rGO due to the efficient restoration of the graphitic structure. We reveal that Ea decreases with the increasing density of states near the Fermi level due to the expansion of the conjugated π-electron system in the rGO. This means that Ea corresponds to the <span class="hlt">energy</span> gap between the top of the valence <span class="hlt">band</span> and the bottom of the conduction <span class="hlt">band</span>. The origin of the <span class="hlt">band</span>-like transport can be explained by the carriers, which are more easily excited into the conduction <span class="hlt">band</span> due to the decreasing <span class="hlt">energy</span> gap with the expansion of the conjugated π-electron system in the rGO.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29082994','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29082994"><span>Investigation of <span class="hlt">energy</span> <span class="hlt">band</span> alignments and interfacial properties of rutile NMO2/TiO2 (NM = Ru, Rh, Os, and Ir) by first-principles calculations.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Yang, Chen; Zhao, Zong-Yan</p> <p>2017-11-08</p> <p>In the field of photocatalysis, constructing hetero-structures is an efficient strategy to improve quantum efficiency. However, a lattice mismatch often induces unfavorable interfacial states that can act as recombination centers for photo-generated electron-hole pairs. If the hetero-structure's components have the same crystal structure, this disadvantage can be easily avoided. Conversely, in the process of loading a noble metal co-catalyst onto the TiO 2 surface, a transition layer of noble metal oxides is often formed between the TiO 2 layer and the noble metal layer. In this article, interfacial properties of hetero-structures composed of a noble metal dioxide and TiO 2 with a rutile crystal structure have been systematically investigated using first-principles calculations. In particular, the Schottky barrier height, <span class="hlt">band</span> bending, and <span class="hlt">energy</span> <span class="hlt">band</span> alignments are studied to provide evidence for practical applications. In all cases, no interfacial states exist in the forbidden <span class="hlt">band</span> of TiO 2 , and the interfacial formation <span class="hlt">energy</span> is very small. A strong internal electric field generated by interfacial electron transfer leads to an efficient separation of photo-generated carriers and <span class="hlt">band</span> bending. Because of the differences in the atomic properties of the components, RuO 2 /TiO 2 and OsO 2 /TiO 2 hetero-structures demonstrate <span class="hlt">band</span> dividing, while RhO 2 /TiO 2 and IrO 2 /TiO 2 hetero-structures have a pseudo-gap near the Fermi <span class="hlt">energy</span> level. Furthermore, NMO 2 /TiO 2 hetero-structures show upward <span class="hlt">band</span> bending. Conversely, RuO 2 /TiO 2 and OsO 2 /TiO 2 hetero-structures present a relatively strong infrared light absorption, while RhO 2 /TiO 2 and IrO 2 /TiO 2 hetero-structures show an obvious absorption edge in the visible light region. Overall, considering all aspects of their properties, RuO 2 /TiO 2 and OsO 2 /TiO 2 hetero-structures are more suitable than others for improving the photocatalytic performance of TiO 2 . These findings will provide useful information</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24329327','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24329327"><span>Passive <span class="hlt">band</span>-gap reconfiguration born from bifurcation asymmetry.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Bernard, Brian P; Mann, Brian P</p> <p>2013-11-01</p> <p>Current periodic structures are constrained to have fixed <span class="hlt">energy</span> transmission behavior unless active control or component replacement is used to alter their wave propagation characteristics. The introduction of nonlinearity to generate multiple stable equilibria is an alternative strategy for realizing distinct <span class="hlt">energy</span> propagation behaviors. We investigate the creation of a reconfigurable <span class="hlt">band</span>-gap system by implementing passive switching between multiple stable states of equilibrium, to alter the level of <span class="hlt">energy</span> attenuation in response to environmental stimuli. The ability to avoid potentially catastrophic loads is demonstrated by tailoring the bandpass and <span class="hlt">band</span>-gap regions to coalesce for two stable equilibria and varying an external load parameter to trigger a bifurcation. The proposed phenomenon could be utilized in remote or autonomous applications where component modifications and active control are impractical.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/486347-table-superdeformed-nuclear-bands-fission-isomers-from-nuclear-data-sheets-issue-may','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/486347-table-superdeformed-nuclear-bands-fission-isomers-from-nuclear-data-sheets-issue-may"><span>Table of superdeformed nuclear <span class="hlt">bands</span> and fission isomers (from Nuclear Data Sheets, v.78, issue 1, May 1996)</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Singh, B.; Firestone, R B.; Chu, S Y.F.</p> <p></p> <p>As part of a committment to maintain nuclear structure data as current as possible in the Evaluated Nuclear Structure Data File (ENSDF) and the Table of Isotopes, the author have been updating the information on superdeformed and hyperdeformed nuclear <span class="hlt">bands</span>. As of February, 1996, they have compiled data for 161 superdeformed <span class="hlt">bands</span> and 47 fission isomers identified in 93 nuclides for this publication. This is an increase of 75 superdeformed <span class="hlt">bands</span> and 20 new nuclides since the first edition in 1994. Partial data for superdeformed <span class="hlt">bands</span> and fission isomers are shown in the <span class="hlt">band</span> drawings. For each nuclide there ismore » a complete level table listing both normal (taken from the ENSDF file) and superdeformed <span class="hlt">band</span> assignments; level <span class="hlt">energy</span>, spin, parity, half-life, magnetic moments, decay branchings; and the <span class="hlt">energies</span>, final levels, relative intensities, multipolarities, and mixing ratios for transitions deexciting each level. Mass excess, decay <span class="hlt">energies</span>, and proton and neutron separation <span class="hlt">energies</span> are also provided from the evaluation of Audi and Wapstra. For superdeformed and hyperdeformed <span class="hlt">bands</span> they provide the following quantities: level <span class="hlt">energies</span>; level half-lives; level spins; and gamma ray <span class="hlt">energies</span>.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1357548-fermi-observations-grb-short-hard-gamma-ray-burst-additional-hard-power-law-component-from-kev-gev-energies','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1357548-fermi-observations-grb-short-hard-gamma-ray-burst-additional-hard-power-law-component-from-kev-gev-energies"><span>Fermi Observations of GRB 090510: A Short Hard Gamma-Ray Burst with an <span class="hlt">Additional</span>, Hard Power-Law Component from 10 keV to GeV <span class="hlt">Energies</span></span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Ackermann, M.; Asano, K.; Atwood, W. B.; ...</p> <p>2010-05-27</p> <p>We present detailed observations of the bright short-hard gamma-ray burst GRB 090510 made with the Gamma-ray Burst Monitor (GBM) and Large Area Telescope (LAT) on board the Fermi observatory. GRB 090510 is the first burst detected by the LAT that shows strong evidence for a deviation from a <span class="hlt">Band</span> spectral fitting function during the prompt emission phase. The time-integrated spectrum is fit by the sum of a <span class="hlt">Band</span> function with E peak = 3.9 ± 0.3 MeV, which is the highest yet measured, and a hard power-law component with photon index –1.62 ± 0.03 that dominates the emission below ≈20more » keV and above ≈100 MeV. The onset of the high-<span class="hlt">energy</span> spectral component appears to be delayed by ~0.1 s with respect to the onset of a component well fit with a single <span class="hlt">Band</span> function. A faint GBM pulse and a LAT photon are detected 0.5 s before the main pulse. During the prompt phase, the LAT detected a photon with <span class="hlt">energy</span> 30.5 +5.8 –2.6 GeV, the highest ever measured from a short GRB. Observation of this photon sets a minimum bulk outflow Lorentz factor, Γ≳ 1200, using simple γγ opacity arguments for this GRB at redshift z = 0.903 and a variability timescale on the order of tens of ms for the ≈100 keV-few MeV flux. Stricter high confidence estimates imply Γ ≳ 1000 and still require that the outflows powering short GRBs are at least as highly relativistic as those of long-duration GRBs. Finally, implications of the temporal behavior and power-law shape of the <span class="hlt">additional</span> component on synchrotron/synchrotron self-Compton, external-shock synchrotron, and hadronic models are considered.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/FR-2013-07-17/pdf/2013-17058.pdf','FEDREG'); return false;" href="https://www.gpo.gov/fdsys/pkg/FR-2013-07-17/pdf/2013-17058.pdf"><span>78 FR 42701 - Improving Public Safety Communications in the 800 MHz <span class="hlt">Band</span></span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collection.action?collectionCode=FR">Federal Register 2010, 2011, 2012, 2013, 2014</a></p> <p></p> <p>2013-07-17</p> <p>...] Improving Public Safety Communications in the 800 MHz <span class="hlt">Band</span> AGENCY: Federal Communications Commission. ACTION...-901 MHz/935- 940 MHz <span class="hlt">band</span> (900 MHz B/ILT <span class="hlt">Band</span>) to allow a qualified entity to file an application for..., manufacturing, <span class="hlt">energy</span>) to non-commercial (e.g., clerical, educational, philanthropic, medical). In 2004, the...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017PhyE...88..142R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017PhyE...88..142R"><span>Multi-shell spherical GaAs /AlxGa1-x As quantum dot shells-size distribution as a mechanism to generate intermediate <span class="hlt">band</span> <span class="hlt">energy</span> levels</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Rodríguez-Magdaleno, K. A.; Pérez-Álvarez, R.; Martínez-Orozco, J. C.; Pernas-Salomón, R.</p> <p>2017-04-01</p> <p>In this work the generation of an intermediate <span class="hlt">band</span> of <span class="hlt">energy</span> levels from multi-shell spherical GaAs /AlxGa1-x As quantum dot shells-size distribution is reported. Within the effective mass approximation the electronic structure of a GaAs spherical quantum-dot surrounded by one, two and three shells is studied in detail using a numerically stable transfer matrix method. We found that a shells-size distribution characterized by continuously wider GaAs domains is a suitable mechanism to generate the intermediate <span class="hlt">band</span> whose width is also dependent on the Aluminium concentration x. Our results suggest that this effective mechanism can be used for the design of wider intermediate <span class="hlt">band</span> than reported in other quantum systems with possible solar cells enhanced performance.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012PhRvB..86s5315L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012PhRvB..86s5315L"><span>Temperature-dependent internal photoemission probe for <span class="hlt">band</span> parameters</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lao, Yan-Feng; Perera, A. G. Unil</p> <p>2012-11-01</p> <p>The temperature-dependent characteristic of <span class="hlt">band</span> offsets at the heterojunction interface was studied by an internal photoemission (IPE) method. In contrast to the traditional Fowler method independent of the temperature (T), this method takes into account carrier thermalization and carrier/dopant-induced <span class="hlt">band</span>-renormalization and <span class="hlt">band</span>-tailing effects, and thus measures the <span class="hlt">band</span>-offset parameter at different temperatures. Despite intensive studies in the past few decades, the T dependence of this key <span class="hlt">band</span> parameter is still not well understood. Re-examining a p-type doped GaAs emitter/undoped AlxGa1-xAs barrier heterojunction system disclosed its previously ignored T dependency in the valence-<span class="hlt">band</span> offset, with a variation up to ˜-10-4 eV/K in order to accommodate the difference in the T-dependent <span class="hlt">band</span> gaps between GaAs and AlGaAs. Through determining the Fermi <span class="hlt">energy</span> level (Ef), IPE is able to distinguish the impurity (IB) and valence <span class="hlt">bands</span> (VB) of extrinsic semiconductors. One important example is to determine Ef of dilute magnetic semiconductors such as GaMnAs, and to understand whether it is in the IB or VB.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017PhyE...85..253D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017PhyE...85..253D"><span>Sizable <span class="hlt">band</span> gap in organometallic topological insulator</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Derakhshan, V.; Ketabi, S. A.</p> <p>2017-01-01</p> <p>Based on first principle calculation when Ceperley-Alder and Perdew-Burke-Ernzerh type exchange-correlation <span class="hlt">energy</span> functional were adopted to LSDA and GGA calculation, electronic properties of organometallic honeycomb lattice as a two-dimensional topological insulator was calculated. In the presence of spin-orbit interaction bulk <span class="hlt">band</span> gap of organometallic lattice with heavy metals such as Au, Hg, Pt and Tl atoms were investigated. Our results show that the organometallic topological insulator which is made of Mercury atom shows the wide bulk <span class="hlt">band</span> gap of about ∼120 meV. Moreover, by fitting the conduction and valence <span class="hlt">bands</span> to the <span class="hlt">band</span>-structure which are produced by Density Functional Theory, spin-orbit interaction parameters were extracted. Based on calculated parameters, gapless edge states within bulk insulating gap are indeed found for finite width strip of two-dimensional organometallic topological insulators.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011PhRvB..84o5308O','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011PhRvB..84o5308O"><span>Modification of the <span class="hlt">band</span> offset in boronitrene</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Obodo, K. O.; Andrew, R. C.; Chetty, N.</p> <p>2011-10-01</p> <p>Using density functional methods within the generalized gradient approximation implemented in the Quantum Espresso codes, we modify the <span class="hlt">band</span> offset in a single layer of boronitrene by substituting a double line of carbon atoms. This effectively introduces a line of dipoles at the interface. We considered various junctions of this system within the zigzag and armchair orientations. Our results show that the “zigzag-short” structure is energetically most stable, with a formation <span class="hlt">energy</span> of 0.502 eV and with a <span class="hlt">band</span> offset of 1.51 eV. The “zigzag-long” structure has a <span class="hlt">band</span> offset of 1.99 eV. The armchair structures are nonpolar, while the zigzag-single structures show a charge accumulation for the C-substituted B and charge depletion for the C-substituted N at the junction. Consequently there is no shifting of the <span class="hlt">bands</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JPCS..115..322Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JPCS..115..322Z"><span>A model for the <span class="hlt">energy</span> <span class="hlt">band</span> gap of GaSbxAs1-x and InSbxAs1-x in the whole composition range</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zhao, Chuan-Zhen; Ren, He-Yu; Wei, Tong; Wang, Sha-Sha; Wang, Jun</p> <p>2018-04-01</p> <p>The <span class="hlt">band</span> gap evolutions of GaSbxAs1-x and InSbxAs1-x in the whole composition range are investigated. It is found that the <span class="hlt">band</span> gap evolutions of GaSbxAs1-x and InSbxAs1-x are determined by two factors. One is the impurity-host interaction in the As-rich and Sb-rich composition ranges. The other is the intraband coupling within the conduction <span class="hlt">band</span> and separately within the valence <span class="hlt">band</span> in the moderate composition range. Based on the <span class="hlt">band</span> gap evolutions of GaSbxAs1-x and InSbxAs1-x, a model is established. In <span class="hlt">addition</span>, it is found that the impurity-host interaction is determined by not only the mismatches in size and electronegativity between the introduced atoms in the host material and the anions of the host material, but also the difference in electronegativity between the introduced atoms in the host material and the cations of the host material.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1984hac..rept.....M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1984hac..rept.....M"><span>Ku-<span class="hlt">Band</span> rendezvous radar performance computer simulation model</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Magnusson, H. G.; Goff, M. F.</p> <p>1984-06-01</p> <p>All work performed on the Ku-<span class="hlt">band</span> rendezvous radar performance computer simulation model program since the release of the preliminary final report is summarized. Developments on the program fall into three distinct categories: (1) modifications to the existing Ku-<span class="hlt">band</span> radar tracking performance computer model; (2) the <span class="hlt">addition</span> of a highly accurate, nonrealtime search and acquisition performance computer model to the total software package developed on this program; and (3) development of radar cross section (RCS) computation models for three <span class="hlt">additional</span> satellites. All changes in the tracking model involved improvements in the automatic gain control (AGC) and the radar signal strength (RSS) computer models. Although the search and acquisition computer models were developed under the auspices of the Hughes Aircraft Company Ku-<span class="hlt">Band</span> Integrated Radar and Communications Subsystem program office, they have been supplied to NASA as part of the Ku-<span class="hlt">band</span> radar performance comuter model package. Their purpose is to predict Ku-<span class="hlt">band</span> acquisition performance for specific satellite targets on specific missions. The RCS models were developed for three satellites: the Long Duration Exposure Facility (LDEF) spacecraft, the Solar Maximum Mission (SMM) spacecraft, and the Space Telescopes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19840024579','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19840024579"><span>Ku-<span class="hlt">Band</span> rendezvous radar performance computer simulation model</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Magnusson, H. G.; Goff, M. F.</p> <p>1984-01-01</p> <p>All work performed on the Ku-<span class="hlt">band</span> rendezvous radar performance computer simulation model program since the release of the preliminary final report is summarized. Developments on the program fall into three distinct categories: (1) modifications to the existing Ku-<span class="hlt">band</span> radar tracking performance computer model; (2) the <span class="hlt">addition</span> of a highly accurate, nonrealtime search and acquisition performance computer model to the total software package developed on this program; and (3) development of radar cross section (RCS) computation models for three <span class="hlt">additional</span> satellites. All changes in the tracking model involved improvements in the automatic gain control (AGC) and the radar signal strength (RSS) computer models. Although the search and acquisition computer models were developed under the auspices of the Hughes Aircraft Company Ku-<span class="hlt">Band</span> Integrated Radar and Communications Subsystem program office, they have been supplied to NASA as part of the Ku-<span class="hlt">band</span> radar performance comuter model package. Their purpose is to predict Ku-<span class="hlt">band</span> acquisition performance for specific satellite targets on specific missions. The RCS models were developed for three satellites: the Long Duration Exposure Facility (LDEF) spacecraft, the Solar Maximum Mission (SMM) spacecraft, and the Space Telescopes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/FR-2012-09-27/pdf/2012-23808.pdf','FEDREG'); return false;" href="https://www.gpo.gov/fdsys/pkg/FR-2012-09-27/pdf/2012-23808.pdf"><span>77 FR 59393 - Jordan Cove <span class="hlt">Energy</span> Project LP; Pacific Connector Gas Pipeline LP; Notice of <span class="hlt">Additional</span> Public...</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collection.action?collectionCode=FR">Federal Register 2010, 2011, 2012, 2013, 2014</a></p> <p></p> <p>2012-09-27</p> <p>...-17-000] Jordan Cove <span class="hlt">Energy</span> Project LP; Pacific Connector Gas Pipeline LP; Notice of <span class="hlt">Additional</span> Public..., and 11, 2012, the Federal <span class="hlt">Energy</span> Regulatory Commission (FERC or Commission) Office of <span class="hlt">Energy</span> Projects... <span class="hlt">additional</span> public scoping meetings to take comments on Jordan Cove <span class="hlt">Energy</span> Project LP's (Jordan Cove) proposed...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24340411','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24340411"><span>Decreasing patient identification <span class="hlt">band</span> errors by standardizing processes.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Walley, Susan Chu; Berger, Stephanie; Harris, Yolanda; Gallizzi, Gina; Hayes, Leslie</p> <p>2013-04-01</p> <p>Patient identification (ID) <span class="hlt">bands</span> are an essential component in patient ID. Quality improvement methodology has been applied as a model to reduce ID <span class="hlt">band</span> errors although previous studies have not addressed standardization of ID <span class="hlt">bands</span>. Our specific aim was to decrease ID <span class="hlt">band</span> errors by 50% in a 12-month period. The Six Sigma DMAIC (define, measure, analyze, improve, and control) quality improvement model was the framework for this study. ID <span class="hlt">bands</span> at a tertiary care pediatric hospital were audited from January 2011 to January 2012 with continued audits to June 2012 to confirm the new process was in control. After analysis, the major improvement strategy implemented was standardization of styles of ID <span class="hlt">bands</span> and labels. <span class="hlt">Additional</span> interventions included educational initiatives regarding the new ID <span class="hlt">band</span> processes and disseminating institutional and nursing unit data. A total of 4556 ID <span class="hlt">bands</span> were audited with a preimprovement ID <span class="hlt">band</span> error average rate of 9.2%. Significant variation in the ID <span class="hlt">band</span> process was observed, including styles of ID <span class="hlt">bands</span>. Interventions were focused on standardization of the ID <span class="hlt">band</span> and labels. The ID <span class="hlt">band</span> error rate improved to 5.2% in 9 months (95% confidence interval: 2.5-5.5; P < .001) and was maintained for 8 months. Standardization of ID <span class="hlt">bands</span> and labels in conjunction with other interventions resulted in a statistical decrease in ID <span class="hlt">band</span> error rates. This decrease in ID <span class="hlt">band</span> error rates was maintained over the subsequent 8 months.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22310942-band-gap-engineering-graphene-using-na-sup-ions','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22310942-band-gap-engineering-graphene-using-na-sup-ions"><span><span class="hlt">Band</span> gap engineering for graphene by using Na{sup +} ions</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Sung, S. J.; Lee, P. R.; Kim, J. G.</p> <p>2014-08-25</p> <p>Despite the noble electronic properties of graphene, its industrial application has been hindered mainly by the absence of a stable means of producing a <span class="hlt">band</span> gap at the Dirac point (DP). We report a new route to open a <span class="hlt">band</span> gap (E{sub g}) at DP in a controlled way by depositing positively charged Na{sup +} ions on single layer graphene formed on 6H-SiC(0001) surface. The doping of low <span class="hlt">energy</span> Na{sup +} ions is found to deplete the π* <span class="hlt">band</span> of graphene above the DP, and simultaneously shift the DP downward away from Fermi <span class="hlt">energy</span> indicating the opening of E{sub g}.more » The <span class="hlt">band</span> gap increases with increasing Na{sup +} coverage with a maximum E{sub g}≥0.70 eV. Our core-level data, C 1s, Na 2p, and Si 2p, consistently suggest that Na{sup +} ions do not intercalate through graphene, but produce a significant charge asymmetry among the carbon atoms of graphene to cause the opening of a <span class="hlt">band</span> gap. We thus provide a reliable way of producing and tuning the <span class="hlt">band</span> gap of graphene by using Na{sup +} ions, which may play a vital role in utilizing graphene in future nano-electronic devices.« less</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li class="active"><span>15</span></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_15 --> <div id="page_16" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li class="active"><span>16</span></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="301"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014APS..MARM37004S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014APS..MARM37004S"><span>Momentum-Space Imaging of the Dirac <span class="hlt">Band</span> Structure in Molecular Graphene via Quasiparticle Interference</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Stephenson, Anna; Gomes, Kenjiro K.; Ko, Wonhee; Mar, Warren; Manoharan, Hari C.</p> <p>2014-03-01</p> <p>Molecular graphene is a nanoscale artificial lattice composed of carbon monoxide molecules arranged one by one, realizing a dream of exploring exotic quantum materials by design. This assembly is done by atomic manipulation with a scanning tunneling microscope (STM) on a Cu(111) surface. To directly probe the transformation of normal surface state electrons into massless Dirac fermions, we map the momentum space dispersion through the Fourier analysis of quasiparticle scattering maps acquired at different <span class="hlt">energies</span> with the STM. The Fourier analysis not only bridges the real-space and momentum-space data but also reveals the chiral nature of those quasiparticles, through a set of selection rules of allowed scattering involving the pseudospin and valley degrees of freedom. The graphene-like <span class="hlt">band</span> structure can be reshaped with simple alterations to the lattice, such as the <span class="hlt">addition</span> of a strain. We analyze the effect on the momentum space <span class="hlt">band</span> structure of multiple types of strain on our system. Supported by DOE, Office of Basic <span class="hlt">Energy</span> Sciences, Division of Materials Sciences and Engineering under contract DE-AC02-76SF00515.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018ascl.soft03003M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018ascl.soft03003M"><span>scarlet: Source separation in multi-<span class="hlt">band</span> images by Constrained Matrix Factorization</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Melchior, Peter; Moolekamp, Fred; Jerdee, Maximilian; Armstrong, Robert; Sun, Ai-Lei; Bosch, James; Lupton, Robert</p> <p>2018-03-01</p> <p>SCARLET performs source separation (aka "deblending") on multi-<span class="hlt">band</span> images. It is geared towards optical astronomy, where scenes are composed of stars and galaxies, but it is straightforward to apply it to other imaging data. Separation is achieved through a constrained matrix factorization, which models each source with a Spectral <span class="hlt">Energy</span> Distribution (SED) and a non-parametric morphology, or multiple such components per source. The code performs forced photometry (with PSF matching if needed) using an optimal weight function given by the signal-to-noise weighted morphology across <span class="hlt">bands</span>. The approach works well if the sources in the scene have different colors and can be further strengthened by imposing various <span class="hlt">additional</span> constraints/priors on each source. Because of its generic utility, this package provides a stand-alone implementation that contains the core components of the source separation algorithm. However, the development of this package is part of the LSST Science Pipeline; the meas_deblender package contains a wrapper to implement the algorithms here for the LSST stack.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/145378','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/145378"><span>Silicone rubber <span class="hlt">band</span> for laparoscopic tubal sterilization.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Ansari, A H; Sealey, R M; Gay, J W; Kang, I</p> <p>1977-12-01</p> <p>In 1974, Yoon and associates (Am J Obstet Gynecol 120:132, 1974) described a new approach in which laparoscopic tubal occlusion was accomplished by utilizing the silicone rubber <span class="hlt">band</span> technique. Recognizing the great advantages of the new technique in eliminating potential thermal injury associated with electrocoagulation, the authors have utilized the Yoon silicone rubber <span class="hlt">band</span> technique in these institutions over the past 20 months. Thus far the procedure has been performed in 304 patients without any major complications. In the hope of eliminating and/or reducing possible pregnancy-failure rates, in 110 cases. In <span class="hlt">addition</span> to application of the silicone <span class="hlt">band</span>, the tube within the <span class="hlt">band</span> was transected with non-electrical Seigler biopsy forceps. This, we believe, should provide an interesting long-term comparative study.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29666394','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29666394"><span>Resonantly enhanced multiple exciton generation through below-<span class="hlt">band</span>-gap multi-photon absorption in perovskite nanocrystals.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Manzi, Aurora; Tong, Yu; Feucht, Julius; Yao, En-Ping; Polavarapu, Lakshminarayana; Urban, Alexander S; Feldmann, Jochen</p> <p>2018-04-17</p> <p>Multi-photon absorption and multiple exciton generation represent two separate strategies for enhancing the conversion efficiency of light into usable electric power. Targeting below-<span class="hlt">band</span>-gap and above-<span class="hlt">band</span>-gap <span class="hlt">energies</span>, respectively, to date these processes have only been demonstrated independently. Here we report the combined interaction of both nonlinear processes in CsPbBr 3 perovskite nanocrystals. We demonstrate nonlinear absorption over a wide range of below-<span class="hlt">band</span>-gap excitation <span class="hlt">energies</span> (0.5-0.8 E g ). Interestingly, we discover high-order absorption processes, deviating from the typical two-photon absorption, at specific energetic positions. These <span class="hlt">energies</span> are associated with a strong enhancement of the photoluminescence intensity by up to 10 5 . The analysis of the corresponding <span class="hlt">energy</span> levels reveals that the observed phenomena can be ascribed to the resonant creation of multiple excitons via the absorption of multiple below-<span class="hlt">band</span>-gap photons. This effect may open new pathways for the efficient conversion of optical <span class="hlt">energy</span>, potentially also in other semiconducting materials.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018InJPh..92..303C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018InJPh..92..303C"><span>A note on anomalous <span class="hlt">band</span>-gap variations in semiconductors with temperature</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Chakraborty, P. K.; Mondal, B. N.</p> <p>2018-03-01</p> <p>An attempt is made to theoretically study the <span class="hlt">band</span>-gap variations (ΔEg) in semiconductors with temperature following the works, did by Fan and O'Donnell et al. based on thermodynamic functions. The semiconductor <span class="hlt">band</span>-gap reflects the bonding <span class="hlt">energy</span>. An increase in temperature changes the chemical bondings, and electrons are promoted from valence <span class="hlt">band</span> to conduction <span class="hlt">band</span>. In their analyses, they made several approximations with respect to temperature and other fitting parameters leading to real values of <span class="hlt">band</span>-gap variations with linear temperature dependences. In the present communication, we have tried to re-analyse the works, specially did by Fan, and derived an analytical model for ΔEg(T). Because, it was based on the second-order perturbation technique of thermodynamic functions. Our analyses are made without any approximations with respect to temperatures and other fitting parameters mentioned in the text, leading to a complex functions followed by an oscillating nature of the variations of ΔEg. In support of the existence of the oscillating <span class="hlt">energy</span> <span class="hlt">band</span>-gap variations with temperature in a semiconductor, possible physical explanations are provided to justify the experimental observation for various materials.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24690441','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24690441"><span>Local <span class="hlt">band</span> gap measurements by VEELS of thin film solar cells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Keller, Debora; Buecheler, Stephan; Reinhard, Patrick; Pianezzi, Fabian; Pohl, Darius; Surrey, Alexander; Rellinghaus, Bernd; Erni, Rolf; Tiwari, Ayodhya N</p> <p>2014-08-01</p> <p>This work presents a systematic study that evaluates the feasibility and reliability of local <span class="hlt">band</span> gap measurements of Cu(In,Ga)Se2 thin films by valence electron <span class="hlt">energy</span>-loss spectroscopy (VEELS). The compositional gradients across the Cu(In,Ga)Se2 layer cause variations in the <span class="hlt">band</span> gap <span class="hlt">energy</span>, which are experimentally determined using a monochromated scanning transmission electron microscope (STEM). The results reveal the expected <span class="hlt">band</span> gap variation across the Cu(In,Ga)Se2 layer and therefore confirm the feasibility of local <span class="hlt">band</span> gap measurements of Cu(In,Ga)Se2 by VEELS. The precision and accuracy of the results are discussed based on the analysis of individual error sources, which leads to the conclusion that the precision of our measurements is most limited by the acquisition reproducibility, if the signal-to-noise ratio of the spectrum is high enough. Furthermore, we simulate the impact of radiation losses on the measured <span class="hlt">band</span> gap value and propose a thickness-dependent correction. In future work, localized <span class="hlt">band</span> gap variations will be measured on a more localized length scale to investigate, e.g., the influence of chemical inhomogeneities and dopant accumulations at grain boundaries.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017NatPh..13..799W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017NatPh..13..799W"><span>Quasiparticle interference and strong electron-mode coupling in the quasi-one-dimensional <span class="hlt">bands</span> of Sr2RuO4</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wang, Zhenyu; Walkup, Daniel; Derry, Philip; Scaffidi, Thomas; Rak, Melinda; Vig, Sean; Kogar, Anshul; Zeljkovic, Ilija; Husain, Ali; Santos, Luiz H.; Wang, Yuxuan; Damascelli, Andrea; Maeno, Yoshiteru; Abbamonte, Peter; Fradkin, Eduardo; Madhavan, Vidya</p> <p>2017-08-01</p> <p>The single-layered ruthenate Sr2RuO4 is presented as a potential spin-triplet superconductor with an order parameter that may break time-reversal invariance and host half-quantized vortices with Majorana zero modes. Although the actual nature of the superconducting state is still a matter of controversy, it is believed to condense from a metallic state that is well described by a conventional Fermi liquid. In this work we use a combination of Fourier transform scanning tunnelling spectroscopy (FT-STS) and momentum-resolved electron <span class="hlt">energy</span> loss spectroscopy (M-EELS) to probe interaction effects in the normal state of Sr2RuO4. Our high-resolution FT-STS data show signatures of the β-<span class="hlt">band</span> with a distinctly quasi-one-dimensional (1D) character. The <span class="hlt">band</span> dispersion reveals surprisingly strong interaction effects that dramatically renormalize the Fermi velocity, suggesting that the normal state of Sr2RuO4 is that of a `correlated metal' where correlations are strengthened by the quasi-1D nature of the <span class="hlt">bands</span>. In <span class="hlt">addition</span>, kinks at <span class="hlt">energies</span> of approximately 10 meV, 38 meV and 70 meV are observed. By comparing STM and M-EELS data we show that the two higher <span class="hlt">energy</span> features arise from coupling with collective modes. The strong correlation effects and the kinks in the quasi-1D <span class="hlt">bands</span> could provide important information for understanding the superconducting state.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/10757939','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/10757939"><span>A Simple <span class="hlt">Band</span> for Gastric <span class="hlt">Banding</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Broadbent</p> <p>1993-08-01</p> <p>The author has noted that flexible gastric <span class="hlt">bands</span> have occasionally stenosed the gastric stoma or allowed it to dilate. A <span class="hlt">band</span> was developed using a soft outer silicone rubber tube over a holding mechanism made out of a nylon cable tie passed within the silicone tube. This simple, easily applied <span class="hlt">band</span> is rigid, resisting scar contracture and dilatation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018MS%26E..347a2017H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018MS%26E..347a2017H"><span>Spin splitting in <span class="hlt">band</span> structures of BiTeX (X=Cl, Br, I) monolayers</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hvazdouski, D. C.; Baranava, M. S.; Stempitsky, V. R.</p> <p>2018-04-01</p> <p>In systems with breaking of inversion symmetry a perpendicular electric field arises that interacts with the conduction electrons. It may give rise to electron state splitting even without influence of external magnetic field due to the spin-orbital interaction (SOI). Such a removal of the spin degeneracy is called the Rashba effect. Nanostructure with the Rashba effect can be part of a spin transistor. Spin degeneracy can be realized in a channel from a material of this type without <span class="hlt">additive</span> of magnetic ions. Lack of <span class="hlt">additive</span> increases the charge carrier mobility and reliability of the device. Ab initio simulations of BiTeX (X=Cl, Br, I) monolayers have been carried out using VASP wherein implemented DFT method. The study of this structures is of interest because such sort of structures can be used their as spin-orbitronics materials. The crystal parameters of BiTeCl, BiTeBr, BiTeI have been determined by the ionic relaxation and static calculations. It is necessary to note that splitting of <span class="hlt">energy</span> <span class="hlt">bands</span> occurs in case of SOI included. The values of the Rashba coefficient aR (in the range from 6.25 to 10.00 eV·Å) have high magnitudes for spintronics materials. <span class="hlt">Band</span> structure of monolayers structures have ideal Rashba electron gas, i.e. there no other <span class="hlt">energy</span> states near to Fermi level except Rashba states.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011PhRvC..83b4316B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011PhRvC..83b4316B"><span>Experimental study of ΔI=1 <span class="hlt">bands</span> in In111</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Banerjee, P.; Ganguly, S.; Pradhan, M. K.; Sharma, H. P.; Muralithar, S.; Singh, R. P.; Bhowmik, R. K.</p> <p>2011-02-01</p> <p>The two ΔI=1 <span class="hlt">bands</span> in In111, built upon the 3461.0 and 4931.8 keV states, have been studied. The <span class="hlt">bands</span> were populated in the reaction Mo100(F19,α4nγ) at a beam <span class="hlt">energy</span> of 105 MeV. Mean lifetimes of nine states, four in the first and five in the second <span class="hlt">band</span>, have been determined for the first time from Doppler shift attenuation data. The deduced B(M1) rates and their behavior as a function of level spin support the interpretation of these <span class="hlt">bands</span> within the framework of the shears mechanism. The geometrical model of Machiavelli has been used to derive the effective gyromagnetic ratios for the two <span class="hlt">bands</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018PhRvP...9e4036A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PhRvP...9e4036A"><span>Point-Defect Nature of the Ultraviolet Absorption <span class="hlt">Band</span> in AlN</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Alden, D.; Harris, J. S.; Bryan, Z.; Baker, J. N.; Reddy, P.; Mita, S.; Callsen, G.; Hoffmann, A.; Irving, D. L.; Collazo, R.; Sitar, Z.</p> <p>2018-05-01</p> <p>We present an approach where point defects and defect complexes are identified using power-dependent photoluminescence excitation spectroscopy, impurity data from SIMS, and density-functional-theory (DFT)-based calculations accounting for the total charge balance in the crystal. Employing the capabilities of such an experimental computational approach, in this work, the ultraviolet-C absorption <span class="hlt">band</span> at 4.7 eV, as well as the 2.7- and 3.9-eV luminescence <span class="hlt">bands</span> in AlN single crystals grown via physical vapor transport (PVT) are studied in detail. Photoluminescence excitation spectroscopy measurements demonstrate the relationship between the defect luminescent <span class="hlt">bands</span> centered at 3.9 and 2.7 eV to the commonly observed absorption <span class="hlt">band</span> centered at 4.7 eV. Accordingly, the thermodynamic transition <span class="hlt">energy</span> for the absorption <span class="hlt">band</span> at 4.7 eV and the luminescence <span class="hlt">band</span> at 3.9 eV is estimated at 4.2 eV, in agreement with the thermodynamic transition <span class="hlt">energy</span> for the CN- point defect. Finally, the 2.7-eV PL <span class="hlt">band</span> is the result of a donor-acceptor pair transition between the VN and CN point defects since nitrogen vacancies are predicted to be present in the crystal in concentrations similar to carbon-employing charge-balance-constrained DFT calculations. Power-dependent photoluminescence measurements reveal the presence of the deep donor state with a thermodynamic transition <span class="hlt">energy</span> of 5.0 eV, which we hypothesize to be nitrogen vacancies in agreement with predictions based on theory. The charge state, concentration, and type of impurities in the crystal are calculated considering a fixed amount of impurities and using a DFT-based defect solver, which considers their respective formation <span class="hlt">energies</span> and the total charge balance in the crystal. The presented results show that nitrogen vacancies are the most likely candidate for the deep donor state involved in the donor-acceptor pair transition with peak emission at 2.7 eV for the conditions relevant to PVT growth.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017A%26A...600A.129E','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017A%26A...600A.129E"><span>Near-infrared diffuse interstellar <span class="hlt">bands</span> in APOGEE telluric standard star spectra . Weak <span class="hlt">bands</span> and comparisons with optical counterparts</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Elyajouri, M.; Lallement, R.; Monreal-Ibero, A.; Capitanio, L.; Cox, N. L. J.</p> <p>2017-04-01</p> <p>Aims: Information on the existence and properties of diffuse interstellar <span class="hlt">bands</span> (DIBs) outside the optical domain is still limited. <span class="hlt">Additional</span> infra-red (IR) measurements and IR-optical correlative studies are needed to constrain DIB carriers and locate various absorbers in 3D maps of the interstellar matter. Methods: We extended our study of H-<span class="hlt">band</span> DIBs in Apache Point Observatory Galactic Evolution Experiment (APOGEE) Telluric Standard Star (TSS) spectra. We used the strong λ15273 <span class="hlt">band</span> to select the most and least absorbed targets. We used individual spectra of the former subsample to extract weaker DIBs, and we searched the two stacked series for differences that could indicate <span class="hlt">additional</span> <span class="hlt">bands</span>. High-resolution NARVAL and SOPHIE optical spectra for a subsample of 55 TSS targets were <span class="hlt">additionally</span> recorded for NIR/optical correlative studies. Results: From the TSS spectra we extract a catalog of measurements of the poorly studied λλ15617, 15653, and 15673 DIBs in ≃300 sightlines, we obtain a first accurate determination of their rest wavelength and constrained their intrinsic width and shape. In <span class="hlt">addition</span>, we studied the relationship between these weak <span class="hlt">bands</span> and the strong λ15273 DIB. We provide a first or second confirmation of several other weak DIBs that have been proposed based on different instruments, and we add new constraints on their widths and locations. We finally propose two new DIB candidates. Conclusions: We compared the strength of the λ15273 absorptions with their optical counterparts λλ5780, 5797, 6196, 6283, and 6614. Using the 5797-5780 ratio as a tracer of shielding against the radiation field, we showed that the λ15273 DIB carrier is significantly more abundant in unshielded (σ-type) clouds, and it responds even more strongly than the λ5780 <span class="hlt">band</span> carrier to the local ionizing field. Full Table 5 is available at the CDS via anonymous ftp to http://cdsarc.u-strasbg.fr (http://130.79.128.5) or via http</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27267558','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27267558"><span>Harvesting Broad Frequency <span class="hlt">Band</span> Blue <span class="hlt">Energy</span> by a Triboelectric-Electromagnetic Hybrid Nanogenerator.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wen, Zhen; Guo, Hengyu; Zi, Yunlong; Yeh, Min-Hsin; Wang, Xin; Deng, Jianan; Wang, Jie; Li, Shengming; Hu, Chenguo; Zhu, Liping; Wang, Zhong Lin</p> <p>2016-07-26</p> <p>Ocean wave associated <span class="hlt">energy</span> is huge, but it has little use toward world <span class="hlt">energy</span>. Although such blue <span class="hlt">energy</span> is capable of meeting all of our <span class="hlt">energy</span> needs, there is no effective way to harvest it due to its low frequency and irregular amplitude, which may restrict the application of traditional power generators. In this work, we report a hybrid nanogenerator that consists of a spiral-interdigitated-electrode triboelectric nanogenerator (S-TENG) and a wrap-around electromagnetic generator (W-EMG) for harvesting ocean <span class="hlt">energy</span>. In this design, the S-TENG can be fully isolated from the external environment through packaging and indirectly driven by the noncontact attractive forces between pairs of magnets, and W-EMG can be easily hybridized. Notably, the hybrid nanogenerator could generate electricity under either rotation mode or fluctuation mode to collect <span class="hlt">energy</span> in ocean tide, current, and wave <span class="hlt">energy</span> due to the unique structural design. In <span class="hlt">addition</span>, the characteristics and advantages of outputs indicate that the S-TENG is irreplaceable for harvesting low rotation speeds (<100 rpm) or motion frequencies (<2 Hz) <span class="hlt">energy</span>, which fits the frequency range for most of the water wave based blue <span class="hlt">energy</span>, while W-EMG is able to produce larger output at high frequencies (>10 Hz). The complementary output can be maximized and hybridized for harvesting <span class="hlt">energy</span> in a broad frequency range. Finally, a single hybrid nanogenerator unit was demonstrated to harvest blue <span class="hlt">energy</span> as a practical power source to drive several LEDs under different simulated water wave conditions. We also proposed a blue <span class="hlt">energy</span> harvesting system floating on the ocean surface that could simultaneously harvest wind, solar, and wave <span class="hlt">energy</span>. The proposed hybrid nanogenerator renders an effective and sustainable progress in practical applications of the hybrid nanogenerator toward harvesting water wave <span class="hlt">energy</span> offered by nature.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012ApPhL.100z3902Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012ApPhL.100z3902Y"><span>Photon ratchet intermediate <span class="hlt">band</span> solar cells</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yoshida, M.; Ekins-Daukes, N. J.; Farrell, D. J.; Phillips, C. C.</p> <p>2012-06-01</p> <p>In this paper, we propose an innovative concept for solar power conversion—the "photon ratchet" intermediate <span class="hlt">band</span> solar cell (IBSC)—which may increase the photovoltaic <span class="hlt">energy</span> conversion efficiency of IBSCs by increasing the lifetime of charge carriers in the intermediate state. The limiting efficiency calculation for this concept shows that the efficiency can be increased by introducing a fast thermal transition of carriers into a non-emissive state. At 1 sun, the introduction of a "ratchet <span class="hlt">band</span>" results in an increase of efficiency from 46.8% to 48.5%, due to suppression of entropy generation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018SuMi..114..169S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018SuMi..114..169S"><span><span class="hlt">Band</span> structure and orbital character of monolayer MoS2 with eleven-<span class="hlt">band</span> tight-binding model</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Shahriari, Majid; Ghalambor Dezfuli, Abdolmohammad; Sabaeian, Mohammad</p> <p>2018-02-01</p> <p>In this paper, based on a tight-binding (TB) model, first we present the calculations of eigenvalues as <span class="hlt">band</span> structure and then present the eigenvectors as probability amplitude for finding electron in atomic orbitals for monolayer MoS2 in the first Brillouin zone. In these calculations we are considering hopping processes between the nearest-neighbor Mo-S, the next nearest-neighbor in-plan Mo-Mo, and the next nearest-neighbor in-plan and out-of-plan S-S atoms in a three-atom based unit cell of two-dimensional rhombic MoS2. The hopping integrals have been solved in terms of Slater-Koster and crystal field parameters. These parameters are calculated by comparing TB model with the density function theory (DFT) in the high-symmetry k-points (i.e. the K- and Γ-points). In our TB model all the 4d Mo orbitals and the 3p S orbitals are considered and detailed analysis of the orbital character of each <span class="hlt">energy</span> level at the main high-symmetry points of the Brillouin zone is described. In comparison with DFT calculations, our results of TB model show a very good agreement for <span class="hlt">bands</span> near the Fermi level. However for other <span class="hlt">bands</span> which are far from the Fermi level, some discrepancies between our TB model and DFT calculations are observed. Upon the accuracy of Slater-Koster and crystal field parameters, on the contrary of DFT, our model provide enough accuracy to calculate all allowed transitions between <span class="hlt">energy</span> <span class="hlt">bands</span> that are very crucial for investigating the linear and nonlinear optical properties of monolayer MoS2.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016PhRvB..94h5403L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016PhRvB..94h5403L"><span>Valley-dependent <span class="hlt">band</span> structure and valley polarization in periodically modulated graphene</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lu, Wei-Tao</p> <p>2016-08-01</p> <p>The valley-dependent <span class="hlt">energy</span> <span class="hlt">band</span> and transport property of graphene under a periodic magnetic-strained field are studied, where the time-reversal symmetry is broken and the valley degeneracy is lifted. The considered superlattice is composed of two different barriers, providing more degrees of freedom for engineering the electronic structure. The electrons near the K and K' valleys are dominated by different effective superlattices. It is found that the <span class="hlt">energy</span> <span class="hlt">bands</span> for both valleys are symmetric with respect to ky=-(AM+ξ AS) /4 under the symmetric superlattices. More finite-<span class="hlt">energy</span> Dirac points, more prominent collimation behavior, and new crossing points are found for K' valley. The degenerate miniband near the K valley splits into two subminibands and produces a new <span class="hlt">band</span> gap under the asymmetric superlattices. The velocity for the K' valley is greatly renormalized compared with the K valley, and so we can achieve a finite velocity for the K valley while the velocity for the K' valley is zero. Especially, the miniband and <span class="hlt">band</span> gap could be manipulated independently, leading to an increase of the conductance. The characteristics of the <span class="hlt">band</span> structure are reflected in the transmission spectra. The Dirac points and the crossing points appear as pronounced peaks in transmission. A remarkable valley polarization is obtained which is robust to the disorder and can be controlled by the strain, the period, and the voltage.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018IJE...105..741S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018IJE...105..741S"><span>Integrated amateur <span class="hlt">band</span> and ultra-wide <span class="hlt">band</span> monopole antenna with multiple <span class="hlt">band</span>-notched</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Srivastava, Kunal; Kumar, Ashwani; Kanaujia, B. K.; Dwari, Santanu</p> <p>2018-05-01</p> <p>This paper presents the integrated amateur <span class="hlt">band</span> and ultra-wide <span class="hlt">band</span> (UWB) monopole antenna with integrated multiple <span class="hlt">band</span>-notched characteristics. It is designed for avoiding the potential interference of frequencies 3.99 GHz (3.83 GHz-4.34 GHz), 4.86 GHz (4.48 GHz-5.63 GHz), 7.20 GHz (6.10 GHz-7.55 GHz) and 8.0 GHz (7.62 GHz-8.47 GHz) with VSWR 4.9, 11.5, 6.4 and 5.3, respectively. Equivalent parallel resonant circuits have been presented for each <span class="hlt">band</span>-notched frequencies of the antenna. Antenna operates in amateur <span class="hlt">band</span> 1.2 GHz (1.05 GHz-1.3 GHz) and UWB <span class="hlt">band</span> from 3.2 GHz-13.9 GHz. Different substrates are used to verify the working of the proposed antenna. Integrated GSM <span class="hlt">band</span> from 0.6 GHz to 1.8 GHz can also be achieved by changing the radius of the radiating patch. Antenna gain varied from 1.4 dBi to 9.8 dBi. Measured results are presented to validate the antenna performances.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011JPhD...44K5101H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011JPhD...44K5101H"><span>Reducing support loss in micromechanical ring resonators using phononic <span class="hlt">band</span>-gap structures</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hsu, Feng-Chia; Hsu, Jin-Chen; Huang, Tsun-Che; Wang, Chin-Hung; Chang, Pin</p> <p>2011-09-01</p> <p>In micromechanical resonators, <span class="hlt">energy</span> loss via supports into the substrates may lead to a low quality factor. To eliminate the support loss, in this paper a phononic <span class="hlt">band</span>-gap structure is employed. We demonstrate a design of phononic-crystal (PC) strips used to support extensional wine-glass mode ring resonators to increase the quality factor. The PC strips are introduced to stop elastic-wave propagation by the <span class="hlt">band</span>-gap and deaf-<span class="hlt">band</span> effects. Analyses of resonant characteristics of the ring resonators and the dispersion relations, eigenmodes, and transmission properties of the PC strips are presented. With the proposed resonator architecture, the finite-element simulations show that the leaky power is effectively reduced and the stored <span class="hlt">energy</span> inside the resonators is enhanced simultaneously as the operating frequencies of the resonators are within the <span class="hlt">band</span> gap or deaf <span class="hlt">bands</span>. Realization of a high quality factor micromechanical ring resonator with minimized support loss is expected.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018NJPh...20d3020A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018NJPh...20d3020A"><span>Dirty two-<span class="hlt">band</span> superconductivity with interband pairing order</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Asano, Yasuhiro; Sasaki, Akihiro; Golubov, Alexander A.</p> <p>2018-04-01</p> <p>We study theoretically the effects of random nonmagnetic impurities on the superconducting transition temperature T c in a two-<span class="hlt">band</span> superconductor characterized by an equal-time s-wave interband pairing order parameter. Because of the two-<span class="hlt">band</span> degree of freedom, it is possible to define a spin-triplet s-wave pairing order parameter as well as a spin-singlet s-wave order parameter. The former belongs to odd-<span class="hlt">band</span>-parity symmetry class, whereas the latter belongs to even-<span class="hlt">band</span>-parity symmetry class. In a spin-singlet superconductor, T c is insensitive to the impurity concentration when we estimate the self-<span class="hlt">energy</span> due to the random impurity potential within the Born approximation. On the other hand in a spin-triplet superconductor, T c decreases with the increase of the impurity concentration. We conclude that Cooper pairs belonging to odd-<span class="hlt">band</span>-parity symmetry class are fragile under the random impurity potential even though they have s-wave pairing symmetry.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JChPh.148t4109F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JChPh.148t4109F"><span><span class="hlt">Band</span> structures in coupled-cluster singles-and-doubles Green's function (GFCCSD)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Furukawa, Yoritaka; Kosugi, Taichi; Nishi, Hirofumi; Matsushita, Yu-ichiro</p> <p>2018-05-01</p> <p>We demonstrate that the coupled-cluster singles-and-doubles Green's function (GFCCSD) method is a powerful and prominent tool drawing the electronic <span class="hlt">band</span> structures and the total <span class="hlt">energies</span>, which many theoretical techniques struggle to reproduce. We have calculated single-electron <span class="hlt">energy</span> spectra via the GFCCSD method for various kinds of systems, ranging from ionic to covalent and van der Waals, for the first time: the one-dimensional LiH chain, one-dimensional C chain, and one-dimensional Be chain. We have found that the bandgap becomes narrower than in HF due to the correlation effect. We also show that the <span class="hlt">band</span> structures obtained from the GFCCSD method include both quasiparticle and satellite peaks successfully. Besides, taking one-dimensional LiH as an example, we discuss the validity of restricting the active space to suppress the computational cost of the GFCCSD method. We show that the calculated results without <span class="hlt">bands</span> that do not contribute to the chemical bonds are in good agreement with full-<span class="hlt">band</span> calculations. With the GFCCSD method, we can calculate the total <span class="hlt">energies</span> and spectral functions for periodic systems in an explicitly correlated manner.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li class="active"><span>16</span></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_16 --> <div id="page_17" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li class="active"><span>17</span></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="321"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25660695','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25660695"><span>The marginal <span class="hlt">band</span> system in nymphalid butterfly wings.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Taira, Wataru; Kinjo, Seira; Otaki, Joji M</p> <p>2015-01-01</p> <p>Butterfly wing color patterns are highly complex and diverse, but they are believed to be derived from the nymphalid groundplan, which is composed of several color pattern systems. Among these pattern systems, the marginal <span class="hlt">band</span> system, including marginal and submarginal <span class="hlt">bands</span>, has rarely been studied. Here, we examined the color pattern diversity of the marginal <span class="hlt">band</span> system among nymphalid butterflies. Marginal and submarginal <span class="hlt">bands</span> are usually expressed as a pair of linear <span class="hlt">bands</span> aligned with the wing margin. However, a submarginal <span class="hlt">band</span> can be expressed as a broken <span class="hlt">band</span>, an elongated oval, or a single dot. The marginal focus, usually a white dot at the middle of a wing compartment along the wing edge, corresponds to the pupal edge spot, one of the pupal cuticle spots that signify the locations of color pattern organizing centers. A marginal <span class="hlt">band</span> can be expressed as a semicircle, an elongated oval, or a pair of eyespot-like structures, which suggest the organizing activity of the marginal focus. Physical damage at the pupal edge spot leads to distal dislocation of the submarginal <span class="hlt">band</span> in Junonia almana and in Vanessa indica, suggesting that the marginal focus functions as an organizing center for the marginal <span class="hlt">band</span> system. Taken together, we conclude that the marginal <span class="hlt">band</span> system is developmentally equivalent to other symmetry systems. <span class="hlt">Additionally</span>, the marginal <span class="hlt">band</span> is likely a core element and the submarginal <span class="hlt">band</span> a paracore element of the marginal <span class="hlt">band</span> system, and both <span class="hlt">bands</span> are primarily specified by the marginal focus organizing center.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28621354','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28621354"><span>Density functional theory calculations for the <span class="hlt">band</span> gap and formation <span class="hlt">energy</span> of Pr4-xCaxSi12O3+xN18-x; a highly disordered compound with low symmetry and a large cell size.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Hong, Sung Un; Singh, Satendra Pal; Pyo, Myoungho; Park, Woon Bae; Sohn, Kee-Sun</p> <p>2017-06-28</p> <p>A novel oxynitride compound, Pr 4-x Ca x Si 12 O 3+x N 18-x , synthesized using a solid-state route has been characterized as a monoclinic structure in the C2 space group using Rietveld refinement on synchrotron powder X-ray diffraction data. The crystal structure of this compound was disordered due to the random distribution of Ca/Pr and N/O ions at various Wyckoff sites. A pragmatic approach for an ab initio calculation based on density function theory (DFT) for this disordered compound has been implemented to calculate an acceptable value of the <span class="hlt">band</span> gap and formation <span class="hlt">energy</span>. In general, for the DFT calculation of a disordered compound, a sufficiently large super cell and infinite variety of ensemble configurations is adopted to simulate the random distribution of ions; however, such an approach is time consuming and cost ineffective. Even a single unit cell model gave rise to 43 008 independent configurations as an input model for the DFT calculations. Since it was nearly impossible to calculate the formation <span class="hlt">energy</span> and the <span class="hlt">band</span> gap <span class="hlt">energy</span> for all 43 008 configurations, an elitist non-dominated sorting genetic algorithm (NSGA-II) was employed to find the plausible configurations. In the NSGA-II, all 43 008 configurations were mathematically treated as genomes and the calculated <span class="hlt">band</span> gap and the formation <span class="hlt">energy</span> as the objective (fitness) function. Generalized gradient approximation (GGA) was first employed in the preliminary screening using NSGA-II, and thereafter a hybrid functional calculation (HSE06) was executed only for the most plausible GGA-relaxed configurations with lower formation and higher <span class="hlt">band</span> gap <span class="hlt">energies</span>. The final <span class="hlt">band</span> gap <span class="hlt">energy</span> (3.62 eV) obtained after averaging over the selected configurations, resembles closely the experimental <span class="hlt">band</span> gap value (4.11 eV).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://eric.ed.gov/?q=music+AND+trends&pg=7&id=EJ898253','ERIC'); return false;" href="https://eric.ed.gov/?q=music+AND+trends&pg=7&id=EJ898253"><span>Steel <span class="hlt">Band</span> Repertoire: The Case for Original Music</span></a></p> <p><a target="_blank" href="http://www.eric.ed.gov/ERICWebPortal/search/extended.jsp?_pageLabel=advanced">ERIC Educational Resources Information Center</a></p> <p>Tanner, Chris</p> <p>2010-01-01</p> <p>In the past few decades, the steel <span class="hlt">band</span> art form has experienced consistent growth and development in several key respects. For example, in the United States, the sheer number of steel <span class="hlt">band</span> programs has steadily increased, and it appears that this trend will continue in the future. <span class="hlt">Additionally</span>, pan builders and tuners have made great strides in…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19790011018','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19790011018"><span>Shuttle Ku-<span class="hlt">band</span> and S-<span class="hlt">band</span> communications implementations study</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Huth, G. K.; Nessibou, T.; Nilsen, P. W.; Simon, M. K.; Weber, C. L.</p> <p>1979-01-01</p> <p>The interfaces between the Ku-<span class="hlt">band</span> system and the TDRSS, between the S-<span class="hlt">band</span> system and the TDRSS, GSTDN and SGLS networks, and between the S-<span class="hlt">band</span> payload communication equipment and the other Orbiter avionic equipment were investigated. The principal activities reported are: (1) performance analysis of the payload narrowband bent-pipe through the Ku-<span class="hlt">band</span> communication system; (2) performance evaluation of the TDRSS user constraints placed on the S-<span class="hlt">band</span> and Ku-<span class="hlt">band</span> communication systems; (3) assessment of the shuttle-unique S-<span class="hlt">band</span> TDRSS ground station false lock susceptibility; (4) development of procedure to make S-<span class="hlt">band</span> antenna measurements during orbital flight; (5) development of procedure to make RFI measurements during orbital flight to assess the performance degradation to the TDRSS S-<span class="hlt">band</span> communication link; and (6) analysis of the payload interface integration problem areas.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013PhyB..422...12B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013PhyB..422...12B"><span>Birefringence and <span class="hlt">band</span> structure of CdP2 crystals</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Beril, S. I.; Stamov, I. G.; Syrbu, N. N.; Zalamai, V. V.</p> <p>2013-08-01</p> <p>The spatial dispersion in CdP2 crystals was investigated. The dispersion is positive (nk||с>nk||у) at λ>λ0 and negative (nk||с<nk||у) at λ<λ0. CdP2 crystals are isotropic for wavelength λо=896 nm. Indirect transitions in excitonic region Еgx are nonpolarized due to one pair of <span class="hlt">bands</span>. Minimal direct <span class="hlt">energy</span> intervals correspond to transitions Г1→Г1 for Е||с and Г2→Г1 for Е⊥с. The temperature coefficient of <span class="hlt">energy</span> gap sifting in the case of temperature changing between 2 and 4.2 K equals to 10.6 meV/K and 3.2 mev/K for Г1→Г1 and Г2→Г1 <span class="hlt">band</span> gap correspondingly. Reflectivity spectra were measured for <span class="hlt">energy</span> interval 1.5-10 eV and optical functions (n, k, ε1, ε2,d2ε1/dE2 and d2ε2/dE2) were calculated by using Kramers-Kronig analyses. All features were interpreted as optical transitions on the basis of both theoretical calculations of <span class="hlt">band</span> structure.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012ApPhL.101h2106G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012ApPhL.101h2106G"><span>Negligible carrier freeze-out facilitated by impurity <span class="hlt">band</span> conduction in highly p-type GaN</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gunning, Brendan; Lowder, Jonathan; Moseley, Michael; Alan Doolittle, W.</p> <p>2012-08-01</p> <p>Highly p-type GaN films with hole concentrations exceeding 6 × 1019 cm-3 grown by metal-modulated epitaxy are electrically characterized. Temperature-dependent Hall effect measurements at cryogenic temperatures reveal minimal carrier freeze-out in highly doped samples, while less heavily doped samples exhibited high resistivity and donor-compensated conductivity as is traditionally observed. Effective activation <span class="hlt">energies</span> as low as 43 meV were extracted, and a maximum Mg activation efficiency of 52% was found. In <span class="hlt">addition</span>, the effective activation <span class="hlt">energy</span> was found to be negatively correlated to the hole concentration. These results indicate the onset of the Mott-Insulator transition leading to impurity <span class="hlt">band</span> conduction.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010AIPC.1299..451F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010AIPC.1299..451F"><span>Ultra-High Gradient S-<span class="hlt">band</span> Linac for Laboratory and Industrial Applications</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Faillace, L.; Agustsson, R.; Dolgashev, V.; Frigola, P.; Murokh, A.; Rosenzweig, J.; Yakimenko, V.</p> <p>2010-11-01</p> <p>A strong demand for high gradient structures arises from the limited real estate available for linear accelerators. RadiaBeam Technologies is developing a Doubled <span class="hlt">Energy</span> Compact Accelerator (DECA) structure: an S-<span class="hlt">band</span> standing wave electron linac designed to operate at accelerating gradients of up to 50 MV/m. In this paper, we present the radio-frequency design of the DECA S-<span class="hlt">band</span> accelerating structure, operating at 2.856 GHz in the π-mode. The structure design is heavily influenced by NLC collaboration experience with ultra high gradient X-<span class="hlt">band</span> structures; S-<span class="hlt">band</span>, however, is chosen to take advantage of commonly available high power S-<span class="hlt">band</span> klystrons.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017ApJ...837..134Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017ApJ...837..134Z"><span>Test of Weak Equivalence Principle with the Multi-<span class="hlt">band</span> Timing of the Crab Pulsar</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zhang, Yueyang; Gong, Biping</p> <p>2017-03-01</p> <p>The Weak Equivalent Principle (WEP) can be tested through the parameterized post-Newtonian parameter γ, representing the space curvature produced by unit rest mass. The parameter γ in turn has been constrained by comparing the arrival times of photons originating in distant transient events, such as gamma-ray bursts, fast radio bursts, and giant pulses of pulsars. Those measurements normally correspond to an individual burst event with very limited <span class="hlt">energy</span> <span class="hlt">bands</span> and signal-to-noise ratios (S/Ns). In this paper, the discrepancy in the pulse arrival times of the Crab Pulsar between different <span class="hlt">energy</span> <span class="hlt">bands</span> is obtained by the phase difference between corresponding pulse profiles. This allows us to compare the pulse arrival times at the largest <span class="hlt">energy</span> <span class="hlt">band</span> differences, between radio and optical, radio and X-ray, and radio and gamma-ray respectively. Because the pulse profiles are generated by phase-folding thousands of individual pulses, the time discrepancies between two <span class="hlt">energy</span> <span class="hlt">bands</span> are actually measured from thousands of events at each <span class="hlt">energy</span> <span class="hlt">band</span>, which corresponds to a much higher S/N. The upper limit of the γ discrepancy set by such an extensively observed and well-modeled source is as follows: {γ }{radio}{--}{γ }γ {- {ray}}< 3.28× {10}-9 at the <span class="hlt">energy</span> difference of {E}γ {- {ray}}/{E}{radio}˜ {10}13, {γ }{radio}{--}{γ }{{X} - {ray}}< 4.01× {10}-9 at the <span class="hlt">energy</span> difference of {E}{{X} - {ray}}/{E}{radio}˜ {10}9, {γ }{radio}{--}{γ }{optical}< 2.63× {10}-9 at {E}{optical}/{E}{radio}˜ {10}5, and {γ }{optical}{--}{γ }γ {- {ray}}< 3.03× {10}-10 at {E}γ {- {ray}}/{E}{optical}˜ {10}8. This actually measures the arrival times of freely falling photons in the gravitational field of the Milky Way with the largest amount of events and with data of the highest S/N, which tests WEP at <span class="hlt">energy</span> <span class="hlt">band</span> differences that have never been reached before.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1212420-development-flexible-free-standing-thin-films-additive-manufacturing-localized-energy-generation','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1212420-development-flexible-free-standing-thin-films-additive-manufacturing-localized-energy-generation"><span>Development of flexible, free-standing, thin films for <span class="hlt">additive</span> manufacturing and localized <span class="hlt">energy</span> generation</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Clark, Billy; McCollum, Jena; Pantoya, Michelle L.; ...</p> <p>2015-08-01</p> <p>Film energetics are becoming increasingly popular because a variety of technologies are driving a need for localized <span class="hlt">energy</span> generation in a stable, safe and flexible form. Aluminum (Al) and molybdenum trioxide (MoO₃) composites were mixed into a silicon binder and extruded using a blade casting technique to form flexible free-standing films ideal for localized <span class="hlt">energy</span> generation. Since this material can be extruded onto a surface it is well suited to <span class="hlt">additive</span> manufacturing applications. This study examines the influence of 0-35% by mass potassium perchlorate (KClO₄) <span class="hlt">additive</span> on the combustion behavior of these energetic films. Without KClO₄ the film exhibits thermalmore » instabilities that produce unsteady <span class="hlt">energy</span> propagation upon reaction. All films were cast at a thickness of 1 mm with constant volume percent solids to ensure consistent rheological properties. The films were ignited and flame propagation was measured. The results show that as the mass percent KClO₄ increased, the flame speed increased and peaked at 0.43 cm/s and 30 wt% KClO₄. Thermochemical equilibrium simulations show that the heat of combustion increases with increasing KClO₄ concentration up to a maximum at 20 wt% when the heat of combustion plateaus, indicating that the increased chemical <span class="hlt">energy</span> liberated by the <span class="hlt">additional</span> KClO₄ promotes stable <span class="hlt">energy</span> propagation. Differential scanning calorimeter and thermogravimetric analysis show that the silicone binder participates as a fuel and reacts with KClO₄ adding <span class="hlt">energy</span> to the reaction and promoting propagation.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19850062559&hterms=poor+performances&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dpoor%2Bperformances','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19850062559&hterms=poor+performances&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dpoor%2Bperformances"><span>Performance of a normalized <span class="hlt">energy</span> metric without jammer state information for an FH/MFSK system in worst case partial <span class="hlt">band</span> jamming</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Lee, P. J.</p> <p>1985-01-01</p> <p>For a frequency-hopped noncoherent MFSK communication system without jammer state information (JSI) in a worst case partial <span class="hlt">band</span> jamming environment, it is well known that the use of a conventional unquantized metric results in very poor performance. In this paper, a 'normalized' unquantized <span class="hlt">energy</span> metric is suggested for such a system. It is shown that with this metric, one can save 2-3 dB in required signal <span class="hlt">energy</span> over the system with hard decision metric without JSI for the same desired performance. When this very robust metric is compared to the conventional unquantized <span class="hlt">energy</span> metric with JSI, the loss in required signal <span class="hlt">energy</span> is shown to be small. Thus, the use of this normalized metric provides performance comparable to systems for which JSI is known. Cutoff rate and bit error rate with dual-k coding are used for the performance measures.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22140248-unidentified-infrared-emission-bands-pahs-maons','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22140248-unidentified-infrared-emission-bands-pahs-maons"><span>UNIDENTIFIED INFRARED EMISSION <span class="hlt">BANDS</span>: PAHs or MAONs?</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Sun Kwok; Yong Zhang, E-mail: sunkwok@hku.hk</p> <p>2013-07-01</p> <p>We suggest that the carrier of the unidentified infrared emission (UIE) <span class="hlt">bands</span> is an amorphous carbonaceous solid with mixed aromatic/aliphatic structures, rather than free-flying polycyclic aromatic hydrocarbon molecules. Through spectral fittings of the astronomical spectra of the UIE <span class="hlt">bands</span>, we show that a significant amount of the <span class="hlt">energy</span> is emitted by the aliphatic component, implying that aliphatic groups are an essential part of the chemical structure. Arguments in favor of an amorphous, solid-state structure rather than a gas-phase molecule as a carrier of the UIE are also presented.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017PhLB..766..107L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017PhLB..766..107L"><span>Observation of a novel stapler <span class="hlt">band</span> in 75As</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Li, C. G.; Chen, Q. B.; Zhang, S. Q.; Xu, C.; Hua, H.; Li, X. Q.; Wu, X. G.; Hu, S. P.; Meng, J.; Xu, F. R.; Liang, W. Y.; Li, Z. H.; Ye, Y. L.; Jiang, D. X.; Sun, J. J.; Han, R.; Niu, C. Y.; Chen, X. C.; Li, P. J.; Wang, C. G.; Wu, H. Y.; Li, G. S.; He, C. Y.; Zheng, Y.; Li, C. B.; Chen, Q. M.; Zhong, J.; Zhou, W. K.</p> <p>2017-03-01</p> <p>The heavy ion fusion-evaporation reaction study for the high-spin spectroscopy of 75As has been performed via the reaction channel 70Zn(9Be, 1p3n)75As at a beam <span class="hlt">energy</span> of 42 MeV. The collective structure especially a dipole <span class="hlt">band</span> in 75As is established for the first time. The properties of this dipole <span class="hlt">band</span> are investigated in terms of the self-consistent tilted axis cranking covariant density functional theory. Based on the theoretical description and the examination of the angular momentum components, this dipole <span class="hlt">band</span> can be interpreted as a novel stapler <span class="hlt">band</span>, where the valence neutrons in (1g9/2) orbital rather than the collective core are responsible for the closing of the stapler of angular momentum.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018AcSpA.196..289R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AcSpA.196..289R"><span>Structural sensitivity of Csbnd H vibrational <span class="hlt">band</span> in methyl benzoate</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Roy, Susmita; Maiti, Kiran Sankar</p> <p>2018-05-01</p> <p>The Csbnd H vibrational <span class="hlt">bands</span> of methyl benzoate are studied to understand its coupling pattern with other vibrational <span class="hlt">bands</span> of the biological molecule. This will facilitate to understand the biological structure and dynamics in spectroscopic as well as in microscopic study. Due to the congested spectroscopic pattern, near degeneracy, and strong anharmonicity of the Csbnd H stretch vibrations, assignment of the Csbnd H vibrational frequencies are often misleading. Anharmonic vibrational frequency calculation with multidimensional potential <span class="hlt">energy</span> surface interprets the Csbnd H vibrational spectra more accurately. In this article we have presented the importance of multidimensional potential <span class="hlt">energy</span> surface in anharmonic vibrational frequency calculation and discuss the unexpected red shift of asymmetric Csbnd H stretch vibration of methyl group. The Csbnd D stretch vibrational <span class="hlt">band</span> which is splitted to double peaks due to the Fermi resonance is also discussed here.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1415918','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1415918"><span>The Current State of <span class="hlt">Additive</span> Manufacturing in Wind <span class="hlt">Energy</span> Systems</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Mann, Margaret; Palmer, Sierra; Lee, Dominic</p> <p></p> <p>Wind power is an inexhaustible form of <span class="hlt">energy</span> that is being captured throughout the U.S. to power the engine of our economy. A robust, domestic wind industry promises to increase U.S. industry growth and competitiveness, strengthen U.S. <span class="hlt">energy</span> security independence, and promote domestic manufacturing nationwide. As of 2016, ~82GW of wind capacity had been installed, and wind power now provides more than 5.5% of the nation’s electricity and supports more than 100,000 domestic jobs, including 500 manufacturing facilities in 43 States. To reach the U.S. Department of Energy’s (DOE’s) 2015 Wind Vision study scenario of wind power serving 35% ofmore » the nation's end-use demand by 2050, significant advances are necessary in all areas of wind technologies and market. An area that can greatly impact the cost and rate of innovation in wind technologies is the use of advanced manufacturing, with one of the most promising areas being <span class="hlt">additive</span> manufacturing (AM). Considering the tremendous promise offered by advanced manufacturing, it is the purpose of this report to identify the use of AM in the production and operation of wind <span class="hlt">energy</span> systems. The report has been produced as a collaborative effort for the DOE Wind <span class="hlt">Energy</span> Technology Office (WETO), between Oak Ridge National Laboratory (ORNL) and the National Renewable <span class="hlt">Energy</span> Laboratory (NREL).« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013PhDT........46G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013PhDT........46G"><span>Redefining RECs: <span class="hlt">Additionality</span> in the voluntary Renewable <span class="hlt">Energy</span> Certificate market</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gillenwater, Michael Wayne</p> <p></p> <p>In the United States, electricity consumers are told that they can "buy" electricity from renewable <span class="hlt">energy</span> projects, versus fossil fuel-fired facilities, through participation in a voluntary green power program. The marketing messages communicate to consumers that their participation and premium payments for a green label will cause <span class="hlt">additional</span> renewable <span class="hlt">energy</span> generation and thereby allow them to claim they consume electricity that is absent pollution as well as reduce pollutant emissions. Renewable <span class="hlt">Energy</span> Certificates (RECs) and wind <span class="hlt">energy</span> are the basis for the majority of the voluntary green power market in the United States. This dissertation addresses the question: Do project developers respond to the voluntary REC market in the United States by altering their decisions to invest in wind turbines? This question is investigated by modeling and probabilistically quantifying the effect of the voluntary REC market on a representative wind power investor in the United States using data from formal expert elicitations of active participants in the industry. It is further explored by comparing the distribution of a sample of wind power projects supplying the voluntary green power market in the United States against an economic viability model that incorporates geographic factors. This dissertation contributes the first quantitative analysis of the effect of the voluntary REC market on project investment. It is found that 1) RECs should be not treated as equivalent to emission offset credits, 2) there is no clearly credible role for voluntary market RECs in emissions trading markets without dramatic restructuring of one or both markets and the environmental commodities they trade, and 3) the use of RECs in entity-level GHG emissions accounting (i.e., "carbon footprinting") leads to double counting of emissions and therefore is not justified. The impotence of the voluntary REC market was, at least in part, due to the small magnitude of the REC price signal and lack of</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29389047','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29389047"><span>Multiscale <span class="hlt">energy</span> reallocation during low-frequency steady-state brain response.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wang, Yifeng; Chen, Wang; Ye, Liangkai; Biswal, Bharat B; Yang, Xuezhi; Zou, Qijun; Yang, Pu; Yang, Qi; Wang, Xinqi; Cui, Qian; Duan, Xujun; Liao, Wei; Chen, Huafu</p> <p>2018-05-01</p> <p>Traditional task-evoked brain activations are based on detection and estimation of signal change from the mean signal. By contrast, the low-frequency steady-state brain response (lfSSBR) reflects frequency-tagging activity at the fundamental frequency of the task presentation and its harmonics. Compared to the activity at these resonant frequencies, brain responses at nonresonant frequencies are largely unknown. <span class="hlt">Additionally</span>, because the lfSSBR is defined by power change, we hypothesize using Parseval's theorem that the power change reflects brain signal variability rather than the change of mean signal. Using a face recognition task, we observed power increase at the fundamental frequency (0.05 Hz) and two harmonics (0.1 and 0.15 Hz) and power decrease within the infra-slow frequency <span class="hlt">band</span> (<0.1 Hz), suggesting a multifrequency <span class="hlt">energy</span> reallocation. The consistency of power and variability was demonstrated by the high correlation (r > .955) of their spatial distribution and brain-behavior relationship at all frequency <span class="hlt">bands</span>. <span class="hlt">Additionally</span>, the reallocation of finite <span class="hlt">energy</span> was observed across various brain regions and frequency <span class="hlt">bands</span>, forming a particular spatiotemporal pattern. Overall, results from this study strongly suggest that frequency-specific power and variability may measure the same underlying brain activity and that these results may shed light on different mechanisms between lfSSBR and brain activation, and spatiotemporal characteristics of <span class="hlt">energy</span> reallocation induced by cognitive tasks. © 2018 Wiley Periodicals, Inc.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016APS..MAR.Y6002G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016APS..MAR.Y6002G"><span>Phonon-induced ultrafast <span class="hlt">band</span> gap control in LaTiO3</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gu, Mingqiang; Rondinelli, James M.</p> <p></p> <p>We propose a route for ultrafast <span class="hlt">band</span> gap engineering in correlated transition metal oxides by using optically driven phonons. We show that the ∖Gamma-point electron <span class="hlt">band</span> <span class="hlt">energies</span> can be deterministically tuned in the nonequilibrium state. Taking the Mott insulator LaTiO3 as an example, we show that such phonon-assisted processes dynamically induce an indirect-to-direct <span class="hlt">band</span> gap transition or even a metal-to-insulator transition, depending on the electron correlation strength. We explain the origin of the dynamical <span class="hlt">band</span> structure control and also establish its generality by examining related oxides. Lastly, we describe experimental routes to realize the <span class="hlt">band</span> structure control with impulsive stimulated Raman scattering.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22492306-development-flexible-free-standing-thin-films-additive-manufacturing-localized-energy-generation','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22492306-development-flexible-free-standing-thin-films-additive-manufacturing-localized-energy-generation"><span>Development of flexible, free-standing, thin films for <span class="hlt">additive</span> manufacturing and localized <span class="hlt">energy</span> generation</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Clark, Billy; McCollum, Jena; Pantoya, Michelle L., E-mail: michelle.pantoya@ttu.edu</p> <p>2015-08-15</p> <p>Film energetics are becoming increasingly popular because a variety of technologies are driving a need for localized <span class="hlt">energy</span> generation in a stable, safe and flexible form. Aluminum (Al) and molybdenum trioxide (MoO{sub 3}) composites were mixed into a silicon binder and extruded using a blade casting technique to form flexible free-standing films ideal for localized <span class="hlt">energy</span> generation. Since this material can be extruded onto a surface it is well suited to <span class="hlt">additive</span> manufacturing applications. This study examines the influence of 0-35% by mass potassium perchlorate (KClO{sub 4}) <span class="hlt">additive</span> on the combustion behavior of these energetic films. Without KClO{sub 4} themore » film exhibits thermal instabilities that produce unsteady <span class="hlt">energy</span> propagation upon reaction. All films were cast at a thickness of 1 mm with constant volume percent solids to ensure consistent rheological properties. The films were ignited and flame propagation was measured. The results show that as the mass percent KClO{sub 4} increased, the flame speed increased and peaked at 0.43 cm/s and 30 wt% KClO{sub 4}. Thermochemical equilibrium simulations show that the heat of combustion increases with increasing KClO{sub 4} concentration up to a maximum at 20 wt% when the heat of combustion plateaus, indicating that the increased chemical <span class="hlt">energy</span> liberated by the <span class="hlt">additional</span> KClO{sub 4} promotes stable <span class="hlt">energy</span> propagation. Differential scanning calorimeter and thermogravimetric analysis show that the silicone binder participates as a fuel and reacts with KClO{sub 4} adding <span class="hlt">energy</span> to the reaction and promoting propagation.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018AIPC.1942i0031T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AIPC.1942i0031T"><span><span class="hlt">Bands</span> dispersion and charge transfer in β-BeH2</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Trivedi, D. K.; Galav, K. L.; Joshi, K. B.</p> <p>2018-04-01</p> <p>Predictive capabilities of ab-initio method are utilised to explore <span class="hlt">bands</span> dispersion and charge transfer in β-BeH2. Investigations are carried out using the linear combination of atomic orbitals method at the level of density functional theory. The crystal structure and related parameters are settled by coupling total <span class="hlt">energy</span> calculations with the Murnaghan equation of state. Electronic <span class="hlt">bands</span> dispersion from PBE-GGA is reported. The PBE-GGA, and PBE0 hybrid functional, show that β-BeH2 is a direct gap semiconductor with 1.18 and 2.40 eV <span class="hlt">band</span> gap. The <span class="hlt">band</span> gap slowly decreases with pressure and beyond l00 GPa overlap of conduction and valence <span class="hlt">bands</span> at the r point is observed. Charge transfer is studied by means of Mullikan population analysis.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22494713-band-band-kp-modeling-electronic-band-structure-material-gain-ga-asbi-quantum-wells-grown-gaas-inp-substrates','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22494713-band-band-kp-modeling-electronic-band-structure-material-gain-ga-asbi-quantum-wells-grown-gaas-inp-substrates"><span>8-<span class="hlt">band</span> and 14-<span class="hlt">band</span> kp modeling of electronic <span class="hlt">band</span> structure and material gain in Ga(In)AsBi quantum wells grown on GaAs and InP substrates</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Gladysiewicz, M.; Wartak, M. S.; Department of Physics and Computer Science, Wilfrid Laurier University, Waterloo, Ontario N2L 3C5</p> <p></p> <p>The electronic <span class="hlt">band</span> structure and material gain have been calculated for GaAsBi/GaAs quantum wells (QWs) with various bismuth concentrations (Bi ≤ 15%) within the 8-<span class="hlt">band</span> and 14-<span class="hlt">band</span> kp models. The 14-<span class="hlt">band</span> kp model was obtained by extending the standard 8-<span class="hlt">band</span> kp Hamiltonian by the valence <span class="hlt">band</span> anticrossing (VBAC) Hamiltonian, which is widely used to describe Bi-related changes in the electronic <span class="hlt">band</span> structure of dilute bismides. It has been shown that in the range of low carrier concentrations n < 5 × 10{sup 18 }cm{sup −3}, material gain spectra calculated within 8- and 14-<span class="hlt">band</span> kp Hamiltonians are similar. It means that the 8-<span class="hlt">band</span> kp model can be usedmore » to calculate material gain in dilute bismides QWs. Therefore, it can be applied to analyze QWs containing new dilute bismides for which the VBAC parameters are unknown. Thus, the <span class="hlt">energy</span> gap and electron effective mass for Bi-containing materials are used instead of VBAC parameters. The electronic <span class="hlt">band</span> structure and material gain have been calculated for 8 nm wide GaInAsBi QWs on GaAs and InP substrates with various compositions. In these QWs, Bi concentration was varied from 0% to 5% and indium concentration was tuned in order to keep the same compressive strain (ε = 2%) in QW region. For GaInAsBi/GaAs QW with 5% Bi, gain peak was determined to be at about 1.5 μm. It means that it can be possible to achieve emission at telecommunication windows (i.e., 1.3 μm and 1.55 μm) for GaAs-based lasers containing GaInAsBi/GaAs QWs. For GaInAsBi/Ga{sub 0.47}In{sub 0.53}As/InP QWs with 5% Bi, gain peak is predicted to be at about 4.0 μm, i.e., at the wavelengths that are not available in current InP-based lasers.« less</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li class="active"><span>17</span></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_17 --> <div id="page_18" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li class="active"><span>18</span></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="341"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/18729418','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/18729418"><span>Carrier multiplication in semiconductor nanocrystals: theoretical screening of candidate materials based on <span class="hlt">band</span>-structure effects.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Luo, Jun-Wei; Franceschetti, Alberto; Zunger, Alex</p> <p>2008-10-01</p> <p>Direct carrier multiplication (DCM) occurs when a highly excited electron-hole pair decays by transferring its excess <span class="hlt">energy</span> to the electrons rather than to the lattice, possibly exciting <span class="hlt">additional</span> electron-hole pairs. Atomistic electronic structure calculations have shown that DCM can be induced by electron-hole Coulomb interactions, in an impact-ionization-like process whose rate is proportional to the density of biexciton states rho XX. Here we introduce a DCM "figure of merit" R2(E) which is proportional to the ratio between the biexciton density of states rhoXX and the single-exciton density of states rhoX, restricted to single-exciton and biexciton states that are coupled by Coulomb interactions. Using R2(E), we consider GaAs, InAs, InP, GaSb, InSb, CdSe, Ge, Si, and PbSe nanocrystals of different sizes. Although DCM can be affected by both quantum-confinement effects (reflecting the underly electronic structure of the confined dot-interior states) and surface effects, here we are interested to isolate the former. To this end the nanocrystal <span class="hlt">energy</span> levels are obtained from the corresponding bulk <span class="hlt">band</span> structure via the truncated crystal approximation. We find that PbSe, Si, GaAs, CdSe, and InP nanocrystals have larger DCM figure of merit than the other nanocrystals. Our calculations suggest that high DCM efficiency requires high degeneracy of the corresponding bulk <span class="hlt">band</span>-edge states. Interestingly, by considering <span class="hlt">band</span> structure effects we find that as the dot size increases the DCM critical <span class="hlt">energy</span> E0 (the <span class="hlt">energy</span> at which R2(E) becomes >or=1) is reduced, suggesting improved DCM. However, whether the normalized E0/epsilong increases or decreases as the dot size increases depends on dot material.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28703225','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28703225"><span>Auger electron emission initiated by the creation of valence-<span class="hlt">band</span> holes in graphene by positron annihilation.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Chirayath, V A; Callewaert, V; Fairchild, A J; Chrysler, M D; Gladen, R W; Mcdonald, A D; Imam, S K; Shastry, K; Koymen, A R; Saniz, R; Barbiellini, B; Rajeshwar, K; Partoens, B; Weiss, A H</p> <p>2017-07-13</p> <p>Auger processes involving the filling of holes in the valence <span class="hlt">band</span> are thought to make important contributions to the low-<span class="hlt">energy</span> photoelectron and secondary electron spectrum from many solids. However, measurements of the <span class="hlt">energy</span> spectrum and the efficiency with which electrons are emitted in this process remain elusive due to a large unrelated background resulting from primary beam-induced secondary electrons. Here, we report the direct measurement of the <span class="hlt">energy</span> spectra of electrons emitted from single layer graphene as a result of the decay of deep holes in the valence <span class="hlt">band</span>. These measurements were made possible by eliminating competing backgrounds by employing low-<span class="hlt">energy</span> positrons (<1.25 eV) to create valence-<span class="hlt">band</span> holes by annihilation. Our experimental results, supported by theoretical calculations, indicate that between 80 and 100% of the deep valence-<span class="hlt">band</span> holes in graphene are filled via an Auger transition.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5511367','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5511367"><span>Auger electron emission initiated by the creation of valence-<span class="hlt">band</span> holes in graphene by positron annihilation</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Chirayath, V. A.; Callewaert, V.; Fairchild, A. J.; Chrysler, M. D.; Gladen, R. W.; Mcdonald, A. D.; Imam, S. K.; Shastry, K.; Koymen, A. R.; Saniz, R.; Barbiellini, B.; Rajeshwar, K.; Partoens, B.; Weiss, A. H.</p> <p>2017-01-01</p> <p>Auger processes involving the filling of holes in the valence <span class="hlt">band</span> are thought to make important contributions to the low-<span class="hlt">energy</span> photoelectron and secondary electron spectrum from many solids. However, measurements of the <span class="hlt">energy</span> spectrum and the efficiency with which electrons are emitted in this process remain elusive due to a large unrelated background resulting from primary beam-induced secondary electrons. Here, we report the direct measurement of the <span class="hlt">energy</span> spectra of electrons emitted from single layer graphene as a result of the decay of deep holes in the valence <span class="hlt">band</span>. These measurements were made possible by eliminating competing backgrounds by employing low-<span class="hlt">energy</span> positrons (<1.25 eV) to create valence-<span class="hlt">band</span> holes by annihilation. Our experimental results, supported by theoretical calculations, indicate that between 80 and 100% of the deep valence-<span class="hlt">band</span> holes in graphene are filled via an Auger transition. PMID:28703225</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1989PhRvB..40.5259W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1989PhRvB..40.5259W"><span>Valence-<span class="hlt">band</span> states in Bi2(Ca,Sr,La)3Cu2O8</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wells, B. O.; Lindberg, P. A. P.; Shen, Z.-X.; Dessau, D. S.; Spicer, W. E.; Lindau, I.; Mitzi, D. B.; Kapitulnik, A.</p> <p>1989-09-01</p> <p>We have used photoemission spectroscopy to examine the symmetry of the occupied states of the valence <span class="hlt">band</span> for the La-doped superconductor Bi2(Ca,Sr,La)3Cu2O8. While the oxygen states near the bottom of the 7-eV wide valence <span class="hlt">band</span> exhibit predominantly O 2pz symmetry, the states at the top of the valence <span class="hlt">band</span> extending to the Fermi level are found to have primarily O 2px and O 2py character. We have also examined anomalous intensity enhancements in the valence-<span class="hlt">band</span> features for photon <span class="hlt">energies</span> near 18 eV. These enhancements, which occur at photon <span class="hlt">energies</span> ranging from 15.8 to 18.0 eV for the different valence-<span class="hlt">band</span> features, are not consistent with either simple final-state effects or direct O 2s transitions to unoccupied O 2p states.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018AtmRe.201..116L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AtmRe.201..116L"><span>Intercomparison of attenuation correction algorithms for single-polarized X-<span class="hlt">band</span> radars</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lengfeld, K.; Berenguer, M.; Sempere Torres, D.</p> <p>2018-03-01</p> <p>Attenuation due to liquid water is one of the largest uncertainties in radar observations. The effects of attenuation are generally inversely proportional to the wavelength, i.e. observations from X-<span class="hlt">band</span> radars are more affected by attenuation than those from C- or S-<span class="hlt">band</span> systems. On the other hand, X-<span class="hlt">band</span> radars can measure precipitation fields in higher temporal and spatial resolution and are more mobile and easier to install due to smaller antennas. A first algorithm for attenuation correction in single-polarized systems was proposed by Hitschfeld and Bordan (1954) (HB), but it gets unstable in case of small errors (e.g. in the radar calibration) and strong attenuation. Therefore, methods have been developed that restrict attenuation correction to keep the algorithm stable, using e.g. surface echoes (for space-borne radars) and mountain returns (for ground radars) as a final value (FV), or adjustment of the radar constant (C) or the coefficient α. In the absence of mountain returns, measurements from C- or S-<span class="hlt">band</span> radars can be used to constrain the correction. All these methods are based on the statistical relation between reflectivity and specific attenuation. Another way to correct for attenuation in X-<span class="hlt">band</span> radar observations is to use <span class="hlt">additional</span> information from less attenuated radar systems, e.g. the ratio between X-<span class="hlt">band</span> and C- or S-<span class="hlt">band</span> radar measurements. Lengfeld et al. (2016) proposed such a method based isotonic regression of the ratio between X- and C-<span class="hlt">band</span> radar observations along the radar beam. This study presents a comparison of the original HB algorithm and three algorithms based on the statistical relation between reflectivity and specific attenuation as well as two methods implementing <span class="hlt">additional</span> information of C-<span class="hlt">band</span> radar measurements. Their performance in two precipitation events (one mainly convective and the other one stratiform) shows that a restriction of the HB is necessary to avoid instabilities. A comparison with vertically pointing</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22280620-fine-structure-red-luminescence-band-undoped-gan','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22280620-fine-structure-red-luminescence-band-undoped-gan"><span>Fine structure of the red luminescence <span class="hlt">band</span> in undoped GaN</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Reshchikov, M. A., E-mail: mreshchi@vcu.edu; Usikov, A.; Saint-Petersburg National Research University of Information Technologies, Mechanics and Optics, 49 Kronverkskiy Ave., 197101 Saint Petersburg</p> <p>2014-01-20</p> <p>Many point defects in GaN responsible for broad photoluminescence (PL) <span class="hlt">bands</span> remain unidentified. Their presence in thick GaN layers grown by hydride vapor phase epitaxy (HVPE) detrimentally affects the material quality and may hinder the use of GaN in high-power electronic devices. One of the main PL <span class="hlt">bands</span> in HVPE-grown GaN is the red luminescence (RL) <span class="hlt">band</span> with a maximum at 1.8 eV. We observed the fine structure of this <span class="hlt">band</span> with a zero-phonon line (ZPL) at 2.36 eV, which may help to identify the related defect. The shift of the ZPL with excitation intensity and the temperature-related transformation of the RLmore » <span class="hlt">band</span> fine structure indicate that the RL <span class="hlt">band</span> is caused by transitions from a shallow donor (at low temperature) or from the conduction <span class="hlt">band</span> (above 50 K) to an unknown deep acceptor having an <span class="hlt">energy</span> level 1.130 eV above the valence <span class="hlt">band</span>.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1999JAP....86.4419P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1999JAP....86.4419P"><span><span class="hlt">Band</span> gap narrowing in n-type and p-type 3C-, 2H-, 4H-, 6H-SiC, and Si</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Persson, C.; Lindefelt, U.; Sernelius, B. E.</p> <p>1999-10-01</p> <p>Doping-induced <span class="hlt">energy</span> shifts of the conduction <span class="hlt">band</span> minimum and the valence <span class="hlt">band</span> maximum have been calculated for n-type and p-type 3C-, 2H-, 4H-, 6H-SiC, and Si. The narrowing of the fundamental <span class="hlt">band</span> gap and of the optical <span class="hlt">band</span> gap are presented as functions of ionized impurity concentration. The calculations go beyond the common parabolic treatments of the ground state <span class="hlt">energy</span> dispersion by using <span class="hlt">energy</span> dispersion and overlap integrals from <span class="hlt">band</span> structure calculations. The nonparabolic valence <span class="hlt">band</span> curvatures influence strongly the <span class="hlt">energy</span> shifts especially in p-type materials. The utilized method is based on a zero-temperature Green's function formalism within the random phase approximation with local field correction according to Hubbard. We have parametrized the shifts of the conduction and the valence <span class="hlt">bands</span> and made comparisons with recently published results from a semi-empirical model.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015JMoSp.313....4U','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015JMoSp.313....4U"><span>Precise ro-vibrational analysis of molecular <span class="hlt">bands</span> forbidden in absorption: The ν8 +ν10 <span class="hlt">band</span> of the 12C2H4 molecule</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ulenikov, O. N.; Gromova, O. V.; Bekhtereva, E. S.; Kashirina, N. V.; Bauerecker, S.; Horneman, V.-M.</p> <p>2015-07-01</p> <p>The highly accurate (experimental accuracy in line positions ∼ (1 - 2) ×10-4 cm-1) ro-vibrational spectrum of the ν8 +ν10 <span class="hlt">band</span> of the 12C2H4 molecule was recorded for the first time with high resolution Fourier transform spectrometry and analyzed in the region of 1650-1950 cm-1 using the Hamiltonian model which takes into account Coriolis resonance interactions between the studied ν8 +ν10 <span class="hlt">band</span>, which is forbidden in absorption, and the <span class="hlt">bands</span> ν4 +ν8 and ν7 +ν8 . About 1570 transitions belonging to the ν8 +ν10 <span class="hlt">band</span> were assigned in the experimental spectra with the maximum values of quantum numbers Jmax. = 35 and Kamax . = 18 . On that basis, a set of 38 vibrational, rotational, centrifugal distortion, and resonance interaction parameters was obtained from the fit. They reproduce values of 598 initial "experimental" ro-vibrational <span class="hlt">energy</span> levels (positions of about 1570 experimentally recorded and assigned transitions) with the rms error drms = 0.00045 cm-1 (drms = 0.00028 cm-1 when upper ro-vibrational <span class="hlt">energies</span> obtained from blended and very weak transitions were deleted from the fit).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28013027','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28013027"><span>Forces directing germ-<span class="hlt">band</span> extension in Drosophila embryos.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kong, Deqing; Wolf, Fred; Großhans, Jörg</p> <p>2017-04-01</p> <p>Body axis elongation by convergent extension is a conserved developmental process found in all metazoans. Drosophila embryonic germ-<span class="hlt">band</span> extension is an important morphogenetic process during embryogenesis, by which the length of the germ-<span class="hlt">band</span> is more than doubled along the anterior-posterior axis. This lengthening is achieved by typical convergent extension, i.e. narrowing the lateral epidermis along the dorsal-ventral axis and simultaneous extension along the anterior-posterior axis. Germ-<span class="hlt">band</span> extension is largely driven by cell intercalation, whose directionality is determined by the planar polarity of the tissue and ultimately by the anterior-posterior patterning system. In <span class="hlt">addition</span>, extrinsic tensile forces originating from the invaginating endoderm induce cell shape changes, which transiently contribute to germ-<span class="hlt">band</span> extension. Here, we review recent progress in understanding of the role of mechanical forces in germ-<span class="hlt">band</span> extension. Copyright © 2016 The Authors. Published by Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20060050766&hterms=Dor&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3DDor','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20060050766&hterms=Dor&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3DDor"><span>Mars Reconnaissance Orbiter Ka-<span class="hlt">band</span> (32 GHz) Demonstration: Cruise Phase Operations</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Shambayati, Shervin; Morabito, David; Border, James S.; Davarian, Faramaz; Lee, Dennis; Mendoza, Ricardo; Britcliffe, Michael; Weinreb, Sander</p> <p>2006-01-01</p> <p>The X-<span class="hlt">band</span> (8.41 GHz) frequency currently used for deep space telecommunications is too narrow (50 MHz) to support future high rate missions. Because of this NASA has decided to transition to Ka-<span class="hlt">band</span> (32 GHz) frequencies. As weather effects cause much larger fluctuations on Ka-<span class="hlt">band</span> than on X-<span class="hlt">band</span>, the traditional method of using a few dBs of margin to cover these fluctuations is wasteful of power for Ka-<span class="hlt">band</span>; therefore, a different operations concept is needed for Ka-<span class="hlt">band</span> links. As part of the development of the operations concept for Ka-<span class="hlt">band</span>, NASA has implemented a fully functioning Ka-<span class="hlt">band</span> communications suite on its Mars Reconnaissance Orbiter (MRO). This suite will be used during the primary science phase to develop and refine the Ka-<span class="hlt">band</span> operations concept for deep space missions. In order to test the functional readiness of the spacecraft and the Deep Space Network's (DSN) readiness to support the demonstration activities a series of passes over DSN 34-m Beam Waveguide (BWG) antennas were scheduled during the cruise phase of the mission. MRO was launched on August 12, 2005 from Kennedy Space Center, Cape Canaveral, Florida, USA and went into Mars Orbit on March 10, 2006. A total of ten telemetry demonstration and one high gain antenna (HGA) calibration passes were allocated to the Ka-<span class="hlt">band</span> demonstration. Furthermore, a number of "shadow" passes were also scheduled where, during a regular MRO track over a Ka-<span class="hlt">band</span> capable antenna, Ka-<span class="hlt">band</span> was identically configured as the X-<span class="hlt">band</span> and tracked by the station. In <span class="hlt">addition</span>, nine Ka-<span class="hlt">band</span> delta differential one way ranging ((delta)DOR) passes were scheduled. During these passes, the spacecraft and the ground system were put through their respective paces. Among the highlights of these was setting a single day record for data return from a deep space spacecraft (133 Gbits) achieved during one 10-hour pass; achieving the highest data rate ever from a planetary mission (6 Mbps) and successfully demonstrating Ka-<span class="hlt">band</span> DDOR</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20170003536&hterms=Currently+Available+Methods+Characterization&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3DCurrently%2BAvailable%2BMethods%2BCharacterization','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20170003536&hterms=Currently+Available+Methods+Characterization&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3DCurrently%2BAvailable%2BMethods%2BCharacterization"><span>Improved <span class="hlt">Band-to-Band</span> Registration Characterization for VIIRS Reflective Solar <span class="hlt">Bands</span> Based on Lunar Observations</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Wang, Zhipeng; Xiong, Xiaoxiong; Li, Yonghong</p> <p>2015-01-01</p> <p>Spectral <span class="hlt">bands</span> of the Visible Infrared Imaging Radiometer Suite (VIIRS) instrumentaboard the Suomi National Polar-orbiting Partnership (S-NPP) satellite are spatially co-registered.The accuracy of the <span class="hlt">band-to-band</span> registration (BBR) is one of the key spatial parameters that must becharacterized. Unlike its predecessor, the Moderate Resolution Imaging Spectroradiometer (MODIS), VIIRS has no on-board calibrator specifically designed to perform on-orbit BBR characterization.To circumvent this problem, a BBR characterization method for VIIRS reflective solar <span class="hlt">bands</span> (RSB) based on regularly-acquired lunar images has been developed. While its results can satisfactorily demonstrate that the long-term stability of the BBR is well within +/- 0.1 moderate resolution bandpixels, undesired seasonal oscillations have been observed in the trending. The oscillations are most obvious between the visiblenear-infrared <span class="hlt">bands</span> and short-middle wave infrared <span class="hlt">bands</span>. This paper investigates the oscillations and identifies their cause as the <span class="hlt">band</span> spectral dependence of the centroid position and the seasonal rotation of the lunar images over calibration events. Accordingly, an improved algorithm is proposed to quantify the rotation and compensate for its impact. After the correction, the seasonal oscillation in the resulting BBR is reduced from up to 0.05 moderate resolution <span class="hlt">band</span> pixels to around 0.01 moderate resolution <span class="hlt">band</span> pixels. After removing this spurious seasonal oscillation, the BBR, as well as its long-term drift are well determined.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2008SPIE.6808E..0AM','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2008SPIE.6808E..0AM"><span>Development of softcopy environment for primary color <span class="hlt">banding</span> visibility assessment</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Min, Byungseok; Pizlo, Zygmunt; Allebach, Jan P.</p> <p>2008-01-01</p> <p>Fine-pitch <span class="hlt">banding</span> is one of the most unwanted artifacts in laser electrophotographic (EP) printers. It is perceived as a quasiperiodic fluctuation in the process direction. Therefore, it is essential for printer vendors to know how <span class="hlt">banding</span> is perceived by humans in order to improve print quality. Monochrome <span class="hlt">banding</span> has been analyzed and assessed by many researchers; but there is no literature that deals with the <span class="hlt">banding</span> of color laser printers as measured from actual prints. The study of color <span class="hlt">banding</span> is complicated by the fact that the color <span class="hlt">banding</span> signal is physically defined in a three-dimensional color space, while <span class="hlt">banding</span> perception is described in a one-dimensional sense such as more <span class="hlt">banding</span> or less <span class="hlt">banding</span>. In <span class="hlt">addition</span>, the color <span class="hlt">banding</span> signal arises from the independent contributions of the four primary colorant <span class="hlt">banding</span> signals. It is not known how these four distinct signals combine to give rise to the perception of color <span class="hlt">banding</span>. In this paper, we develop a methodology to assess the <span class="hlt">banding</span> visibility of the primary colorant cyan based on human visual perception. This is our first step toward studying the more general problem of color <span class="hlt">banding</span> in combinations of two or more colorants. According to our method, we print and scan the cyan test patch, and extract the <span class="hlt">banding</span> profile as a one dimensional signal so that we can freely adjust the intensity of <span class="hlt">banding</span>. Thereafter, by exploiting the pulse width modulation capability of the laser printer, the extracted <span class="hlt">banding</span> profile is used to modulate a pattern consisting of periodic lines oriented in the process direction, to generate extrinsic <span class="hlt">banding</span>. This avoids the effect of the halftoning algorithm on the <span class="hlt">banding</span>. Furthermore, to conduct various <span class="hlt">banding</span> assessments more efficiently, we also develop a softcopy environment that emulates a hardcopy image on a calibrated monitor, which requires highly accurate device calibration throughout the whole system. To achieve the same color appearance as the hardcopy</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://medlineplus.gov/ency/article/007388.htm','NIH-MEDLINEPLUS'); return false;" href="https://medlineplus.gov/ency/article/007388.htm"><span>Laparoscopic gastric <span class="hlt">banding</span></span></a></p> <p><a target="_blank" href="http://medlineplus.gov/">MedlinePlus</a></p> <p></p> <p></p> <p>... adjustable gastric <span class="hlt">banding</span>; Bariatric surgery - laparoscopic gastric <span class="hlt">banding</span>; Obesity - gastric <span class="hlt">banding</span>; Weight loss - gastric <span class="hlt">banding</span> ... gastric <span class="hlt">banding</span> is not a "quick fix" for obesity. It will greatly change your lifestyle. You must ...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016PhLA..380.2836Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016PhLA..380.2836Y"><span>Topologically trivial and nontrivial edge <span class="hlt">bands</span> in graphene induced by irradiation</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yang, Mou; Cai, Zhi-Jun; Wang, Rui-Qiang; Bai, Yan-Kui</p> <p>2016-08-01</p> <p>We proposed a minimal model to describe the Floquet <span class="hlt">band</span> structure of two-dimensional materials with light-induced resonant inter-<span class="hlt">band</span> transition. We applied it to graphene to study the <span class="hlt">band</span> features caused by the light irradiation. Linearly polarized light induces pseudo gaps (gaps are functions of wavevector), and circularly polarized light causes real gaps on the quasi-<span class="hlt">energy</span> spectrum. If the polarization of light is linear and along the longitudinal direction of zigzag ribbons, flat edge <span class="hlt">bands</span> appear in the pseudo gaps, and if it is in the lateral direction of armchair ribbons, curved edge <span class="hlt">bands</span> can be found. For the circularly polarized cases, edge <span class="hlt">bands</span> arise and intersect in the gaps of both types of ribbons. The edge <span class="hlt">bands</span> induced by the circularly polarized light are helical and those by linearly polarized light are topologically trivial ones. The Chern number of the Floquet <span class="hlt">band</span>, which reflects the number of pairs of helical edge <span class="hlt">bands</span> in graphene ribbons, can be reduced into the winding number at resonance.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010OptCo.283.4298M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010OptCo.283.4298M"><span><span class="hlt">Band</span> gaps in periodically magnetized homogeneous anisotropic media</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Merzlikin, A. M.; Levy, M.; Vinogradov, A. P.; Wu, Z.; Jalali, A. A.</p> <p>2010-11-01</p> <p>In [A. M. Merzlikin, A. P. Vinogradov, A. V. Dorofeenko, M. Inoue, M. Levy, A. B. Granovsky, Physica B 394 (2007) 277] it is shown that in anisotropic magnetophotonic crystal made of anisotropic dielectric layers and isotropic magneto-optical layers the magnetization leads to formation of <span class="hlt">additional</span> <span class="hlt">band</span> gaps (BG) inside the Brillouin zones. Due to the weakness of the magneto-optical effects the width of these BG is much smaller than that of usual BG forming on the boundaries of Brillouin zones. In the present communication we show that though the anisotropy suppresses magneto-optical effects. An anisotropic magnetophotonic crystal made of anisotropic dielectric layers and anisotropic magneto-optical; the width of <span class="hlt">additional</span> BG may be much greater than the width of the usual Brillouin BG. Anisotropy tends to suppress Brillouin zone boundary <span class="hlt">band</span> gap formation because the anisotropy suppresses magneto-optical properties, while degenerate <span class="hlt">band</span> gap formation occurs around points of effective isotropy and is not suppressed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015APS..GECGT1091P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015APS..GECGT1091P"><span>Numerical <span class="hlt">band</span> structure calculations of plasma metamaterials</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Pederson, Dylan; Kourtzanidis, Konstantinos; Raja, Laxminarayan</p> <p>2015-09-01</p> <p>Metamaterials (MM) are materials engineered to display negative macroscopic permittivity and permeability. These materials allow for designed control over electromagnetic <span class="hlt">energy</span> flow, especially at frequencies where natural materials do not interact. Plasmas have recently found application in MM as a negative permittivity component. The permittivity of a plasma depends on its electron density, which can be controlled by an applied field. This means that plasmas can be used in MM to actively control the transmission or reflection of incident waves. This work focuses on a plasma MM geometry in which microplasmas are generated in perforations in a metal plate. We characterizethis material by its <span class="hlt">band</span> structure, which describes its interaction with incident waves. The plasma-EM interactions are obtained by coupling Maxwell's equations to a simplified plasma momentum equation. A plasma density profile is prescribed, and its effect on the <span class="hlt">band</span> structure is investigated. The <span class="hlt">band</span> structure calculations are typically done for static structures, whereas our current density responds to the incident waves. The resulting <span class="hlt">band</span> structures are compared with experimental results.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28256467','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28256467"><span>Antibacterial nanosilver coated orthodontic <span class="hlt">bands</span> with potential implications in dentistry.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Prabha, Rahul Damodaran; Kandasamy, Rajasigamani; Sivaraman, U Sajeev; Nandkumar, Maya A; Nair, Prabha D</p> <p>2016-10-01</p> <p>Fixed orthodontic treatment, an indispensable procedure in orthodontics, necessitates insertion of dental <span class="hlt">bands</span>. Insertion of <span class="hlt">band</span> material could also introduce a site of plaque retention. It was hypothesized that <span class="hlt">band</span> materials with slow-release antimicrobial properties could help in sustained infection control, prevention of dental plaque formation and further associated health risks. Considering the known antimicrobial proprieties of silver, a coating of silver nanoparticle (SNP) onto the stainless steel <span class="hlt">bands</span> was done and characterized for its beneficial properties in the prevention of plaque accumulation. Coatings of SNPs on conventional stainless steel dental <span class="hlt">bands</span> were prepared using thermal evaporation technology. The coated dental <span class="hlt">bands</span> were characterized for their physicochemical properties and evaluated for antimicrobial activity and biocompatibility. The physiochemical characterization of <span class="hlt">band</span> material both coated and uncoated was carried out using scanning electron microscope, <span class="hlt">energy</span> dispersive spectroscopy, atomic force microscopyand contact angle test. Biocompatibility tests for coated <span class="hlt">band</span> material were carried using L929 mouse fibroblast cell culture and MTT [3-(4, 5-dimethyl thiazol-2-yl)-2, 5-diphenyl tetrazolium bromide] assay. Antimicrobial activity of coated <span class="hlt">band</span> material against Gram-positive bacteria was tested. A stable and uniform coating of SNPs was obtained. The coated <span class="hlt">band</span> materials were biocompatible as well as possessed distinct antimicrobial activity. The SNP coated dental <span class="hlt">bands</span> could be potential antimicrobial dental <span class="hlt">bands</span> for future clinical use. Further studies need to be done to validate the efficiency of coated <span class="hlt">band</span> materials in oral environments.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018AIPA....8e5317Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AIPA....8e5317Z"><span>Torsional wave <span class="hlt">band</span> gap properties in a circular plate of a two-dimensional generalized phononic crystal</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zhao, Lei; Shu, Haisheng; Liang, Shanjun; Shi, Xiaona; An, Shuowei; Ren, Wanyue; Zhu, Jie</p> <p>2018-05-01</p> <p>The torsional wave <span class="hlt">band</span> gap properties of a two-dimensional generalized phononic crystal (GPC) are investigated in this paper. The GPC structure considered is consisted of two different materials being arranged with radial and circumferential periodicities simultaneously. Based on the viewpoint of <span class="hlt">energy</span> distribution and the finite element method, the power flow, <span class="hlt">energy</span> density, sound intensity vector together with the stress field of the structure excited by torsional load are numerically calculated and discussed. Our results show that, the <span class="hlt">band</span> gap of Bragg type exists in these two-dimensional composite structures, and the <span class="hlt">band</span> gap range is mainly determined by radial periodicity while the circumferential periodicity would result in some transmission peaks within the <span class="hlt">band</span> gap. These peaks are mainly produced by two different mechanisms, the <span class="hlt">energy</span> leakage occurred in circumferential channels and the excitation of the local eigenmodes of certain scatterers. These results may be useful in torsional vibration control for various rotational parts and components, and in the application of <span class="hlt">energy</span> harvesting, etc.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26222731','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26222731"><span>On Valence-<span class="hlt">Band</span> Splitting in Layered MoS2.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zhang, Youwei; Li, Hui; Wang, Haomin; Liu, Ran; Zhang, Shi-Li; Qiu, Zhi-Jun</p> <p>2015-08-25</p> <p>As a representative two-dimensional semiconducting transition-metal dichalcogenide (TMD), the electronic structure in layered MoS2 is a collective result of quantum confinement, interlayer interaction, and crystal symmetry. A prominent <span class="hlt">energy</span> splitting in the valence <span class="hlt">band</span> gives rise to many intriguing electronic, optical, and magnetic phenomena. Despite numerous studies, an experimental determination of valence-<span class="hlt">band</span> splitting in few-layer MoS2 is still lacking. Here, we show how the valence-<span class="hlt">band</span> maximum (VBM) splits for one to five layers of MoS2. Interlayer coupling is found to contribute significantly to phonon <span class="hlt">energy</span> but weakly to VBM splitting in bilayers, due to a small interlayer hopping <span class="hlt">energy</span> for holes. Hence, spin-orbit coupling is still predominant in the splitting. A temperature-independent VBM splitting, known for single-layer MoS2, is, thus, observed for bilayers. However, a Bose-Einstein type of temperature dependence of VBM splitting prevails in three to five layers of MoS2. In such few-layer MoS2, interlayer coupling is enhanced with a reduced interlayer distance, but thermal expansion upon temperature increase tends to decouple adjacent layers and therefore decreases the splitting <span class="hlt">energy</span>. Our findings that shed light on the distinctive behaviors about VBM splitting in layered MoS2 may apply to other hexagonal TMDs as well. They will also be helpful in extending our understanding of the TMD electronic structure for potential applications in electronics and optoelectronics.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017ApSS..396.1562P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017ApSS..396.1562P"><span>Reconstructing the <span class="hlt">energy</span> <span class="hlt">band</span> electronic structure of pulsed laser deposited CZTS thin films intended for solar cell absorber applications</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Pandiyan, Rajesh; Oulad Elhmaidi, Zakaria; Sekkat, Zouheir; Abd-lefdil, Mohammed; El Khakani, My Ali</p> <p>2017-02-01</p> <p>We report here on the use of pulsed KrF-laser deposition (PLD) technique for the growth of high-quality Cu2ZnSnS4 (CZTS) thin films onto Si, and glass substrates without resorting to any post sulfurization process. The PLD-CZTS films were deposited at room temperature (RT) and then subjected to post annealing at different temperatures ranging from 200 to 500 °C in Argon atmosphere. The X-ray diffraction and Raman spectroscopy confirmed that the PLD films crystallize in the characteristic kesterite CZTS structure regardless of their annealing temperature (Ta), but their crystallinity is much improved for Ta ≥ 400 °C. The PLD-CZTS films were found to exhibit a relatively dense morphology with a surface roughness (RMS) that increases with Ta (from ∼14 nm at RT to 70 nm at Ta = 500 °C with a value around 40 nm for Ta = 300-400 °C). The optical bandgap of the PLD-CZTS films, was derived from UV-vis transmission spectra analysis, and found to decrease from 1.73 eV for non-annealed films to ∼1.58 eV for those annealed at Ta = 300 °C. These <span class="hlt">band</span> gap values are very close to the optimum value needed for an ideal solar cell absorber. In order to achieve a complete reconstruction of the one-dimensional <span class="hlt">energy</span> <span class="hlt">band</span> structure of these PLD-CZTS absorbers, we have combined both XPS and UPS spectroscopies to determine their chemical bondings, the position of their valence <span class="hlt">band</span> maximum (relative to Fermi level), and their work function values. This enabled us to sketch out, as accurately as possible, the <span class="hlt">band</span> alignment of the heterojunction interface formed between CZTS and both CdS and ZnS buffer layer materials.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li class="active"><span>18</span></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_18 --> <div id="page_19" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li class="active"><span>19</span></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="361"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/16392585','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/16392585"><span>Correlation effects and electronic properties of Cr-substituted SZn with an intermediate <span class="hlt">band</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Tablero, C</p> <p>2005-09-15</p> <p>A study using first principles of the electronic properties of S32Zn31Cr, a material derived from the SZn host semiconductor where a Cr atom has been substituted for each of the 32 Zn atoms, is presented. This material has an intermediate <span class="hlt">band</span> sandwiched between the valence and conduction <span class="hlt">bands</span> of the host semiconductor, which in a formal <span class="hlt">band</span>-theoretic picture is metallic because the Fermi <span class="hlt">energy</span> is located within the impurity <span class="hlt">band</span>. The potential technological application of these materials is that when they are used to absorb photons in solar cells, the efficiency increases significantly with respect to the host semiconductor. An analysis of the gaps, bandwidths, density of states, total and orbital charges, and electronic density is carried out. The main effects of the local-density approximation with a Hubbard term corrections are an increase in the bandwidth, a modification of the relative composition of the five d and p transition-metal orbitals, and a splitting of the intermediate <span class="hlt">band</span>. The results demonstrate that the main contribution to the intermediate <span class="hlt">band</span> is the Cr atom. For values of U greater than 6 eV, where U is the empirical Hubbard term U parameter, this <span class="hlt">band</span> is unfolded, thus creating two <span class="hlt">bands</span>, a full one below the Fermi <span class="hlt">energy</span> and an empty one above it, i.e., a metal-insulator transition.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19720011057','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19720011057"><span>Excitation of the Werner <span class="hlt">bands</span> of H2 by electron impact</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Stone, E. J.; Zipf, E. C.</p> <p>1972-01-01</p> <p>Absolute cross sections for the excitation of the H2 Werner <span class="hlt">band</span> system were measured from <span class="hlt">energy</span> threshold to 300 eV for electron impact on H2. The <span class="hlt">bands</span> were observed in emission in the wavelength region 1100A to 1250A. The measured cross sections were compared with published transition probabilities, leading to the conclusion that the Werner <span class="hlt">bands</span> are suitable as the basis for a relative spectral response calibration only when the <span class="hlt">bands</span> are observed under sufficiently high resolution. The effect of the perturbation between the C 1Pi u and B 1 Sigma-u states of the hydrogen molecule was clearly observed in anomalies in the rotational intensity distribution in <span class="hlt">bands</span> of the (3 v '') progression.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017APS..MARL33001M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017APS..MARL33001M"><span><span class="hlt">Band</span> alignment in atomically precise graphene nanoribbon junctions</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ma, Chuanxu; Liang, Liangbo; Hong, Kunlun; Li, An-Ping; Xiao, Zhongcan; Lu, Wenchang; Bernholc, Jerry</p> <p></p> <p>Building atomically precise graphene nanoribbon (GNR) heterojunctions down to molecular level opens a new realm to functional graphene-based devices. By employing a surface-assisted self-assembly process, we have synthesized heterojunctions of armchair GNRs (aGNR) with widths of seven, fourteen and twenty-one carbon atoms, denoted 7, 14 and 21-aGNR respectively. A combined study with scanning tunneling microscopy (STM) and density functional theory (DFT) allows the visualization of electronic <span class="hlt">band</span> structures and <span class="hlt">energy</span> level alignments at the heterojunctions with varying widths. A wide bandgap ( 2.6 eV) has been identified on semiconducting 7-aGNR, while the 14-aGNR appears nearly metallic and the 21-aGNR possesses a narrow bandgap. The spatially modulations of the <span class="hlt">energy</span> <span class="hlt">bands</span> are strongly confined at the heterojunctions within a width of about 2 nm. Clear <span class="hlt">band</span> bending of about 0.4 eV and 0.1 eV are observed at the 7-14 and 14-21 aGNR heterojunctions, respectively. This research was conducted at the Center for Nanophase Materials Sciences, which is a DOE Office of Science User Facility.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22566421','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22566421"><span>Reduction in pediatric identification <span class="hlt">band</span> errors: a quality collaborative.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Phillips, Shannon Connor; Saysana, Michele; Worley, Sarah; Hain, Paul D</p> <p>2012-06-01</p> <p>Accurate and consistent placement of a patient identification (ID) <span class="hlt">band</span> is used in health care to reduce errors associated with patient misidentification. Multiple safety organizations have devoted time and <span class="hlt">energy</span> to improving patient ID, but no multicenter improvement collaboratives have shown scalability of previously successful interventions. We hoped to reduce by half the pediatric patient ID <span class="hlt">band</span> error rate, defined as absent, illegible, or inaccurate ID <span class="hlt">band</span>, across a quality improvement learning collaborative of hospitals in 1 year. On the basis of a previously successful single-site intervention, we conducted a self-selected 6-site collaborative to reduce ID <span class="hlt">band</span> errors in heterogeneous pediatric hospital settings. The collaborative had 3 phases: preparatory work and employee survey of current practice and barriers, data collection (ID <span class="hlt">band</span> failure rate), and intervention driven by data and collaborative learning to accelerate change. The collaborative audited 11377 patients for ID <span class="hlt">band</span> errors between September 2009 and September 2010. The ID <span class="hlt">band</span> failure rate decreased from 17% to 4.1% (77% relative reduction). Interventions including education of frontline staff regarding correct ID <span class="hlt">bands</span> as a safety strategy; a change to softer ID <span class="hlt">bands</span>, including "luggage tag" type ID <span class="hlt">bands</span> for some patients; and partnering with families and patients through education were applied at all institutions. Over 13 months, a collaborative of pediatric institutions significantly reduced the ID <span class="hlt">band</span> failure rate. This quality improvement learning collaborative demonstrates that safety improvements tested in a single institution can be disseminated to improve quality of care across large populations of children.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24472000','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24472000"><span><span class="hlt">Energy-band</span> engineering for tunable memory characteristics through controlled doping of reduced graphene oxide.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Han, Su-Ting; Zhou, Ye; Yang, Qing Dan; Zhou, Li; Huang, Long-Biao; Yan, Yan; Lee, Chun-Sing; Roy, Vellaisamy A L</p> <p>2014-02-25</p> <p>Tunable memory characteristics are used in multioperational mode circuits where memory cells with various functionalities are needed in one combined device. It is always a challenge to obtain control over threshold voltage for multimode operation. On this regard, we use a strategy of shifting the work function of reduced graphene oxide (rGO) in a controlled manner through doping gold chloride (AuCl3) and obtained a gradient increase of rGO work function. By inserting doped rGO as floating gate, a controlled threshold voltage (Vth) shift has been achieved in both p- and n-type low voltage flexible memory devices with large memory window (up to 4 times for p-type and 8 times for n-type memory devices) in comparison with pristine rGO floating gate memory devices. By proper <span class="hlt">energy</span> <span class="hlt">band</span> engineering, we demonstrated a flexible floating gate memory device with larger memory window and controlled threshold voltage shifts.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015JPhD...48O5302A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015JPhD...48O5302A"><span>Electrical and <span class="hlt">band</span> structural analyses of Ti1-x Al x O y films grown by atomic layer deposition on p-type GaAs</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>An, Youngseo; Mahata, Chandreswar; Lee, Changmin; Choi, Sungho; Byun, Young-Chul; Kang, Yu-Seon; Lee, Taeyoon; Kim, Jiyoung; Cho, Mann-Ho; Kim, Hyoungsub</p> <p>2015-10-01</p> <p>Amorphous Ti1-x Al x O y films in the Ti-oxide-rich regime (x  <  0.5) were deposited on p-type GaAs via atomic layer deposition with titanium isopropoxide, trimethylaluminum, and H2O precursor chemistry. The electrical properties and <span class="hlt">energy</span> <span class="hlt">band</span> alignments were examined for the resulting materials with their underlying substrates, and significant frequency dispersion was observed in the accumulation region of the Ti-oxide-rich Ti1-x Al x O y films. Although a further reduction in the frequency dispersion and leakage current (under gate electron injection) could be somewhat achieved through a greater <span class="hlt">addition</span> of Al-oxide in the Ti1-x Al x O y film, the simultaneous decrease in the dielectric constant proved problematic in finding an optimal composition for application as a gate dielectric on GaAs. The spectroscopic <span class="hlt">band</span> alignment measurements of the Ti-oxide-rich Ti1-x Al x O y films indicated that the <span class="hlt">band</span> gaps had a rather slow increase with the <span class="hlt">addition</span> of Al-oxide, which was primarily compensated for by an increase in the valance <span class="hlt">band</span> offset, while a nearly-constant conduction <span class="hlt">band</span> offset with a negative electron barrier height was maintained.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014PhyE...63..264M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014PhyE...63..264M"><span><span class="hlt">Band</span> gap opening in α-graphyne by adsorption of organic molecule</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Majidi, R.; Karami, A. R.</p> <p>2014-09-01</p> <p>The lack of a <span class="hlt">band</span> gap limits the application of graphyne in nanoelectronic devices. We have investigated possibility of opening a <span class="hlt">band</span> gap in α-graphyne by adsorption of tetracyanoethylene. The electronic property of α-graphyne in the presence of different numbers of tetracyanoethylene has been studied using density functional theory. It is found that charge is transferred from graphyne sheet to tetracyanoethylene molecules. In the presence of this electron acceptor molecule, a semimetal α-graphyne shows semiconducting property. The <span class="hlt">energy</span> <span class="hlt">band</span> gap at the Dirac point is enhanced by increasing the number of tetracyanoethylene. Our results provide a simple method to create and control the <span class="hlt">band</span> gap in α-graphyne.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2008MMTA...39.1727P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2008MMTA...39.1727P"><span>Study of Inclusion <span class="hlt">Bands</span> in Continuously Cast Steel Billets for Rolling Thermomechanically Treated Rebars</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Pandey, J. C.; Choubey, P. N.; Raj, Manish</p> <p>2008-07-01</p> <p>The article presents the investigation results of inclusion <span class="hlt">bands</span> generally formed toward the loose end/inner radius of continuously cast (CC) strands of thermomechanically treated (TMT) rebar grade cast through curved molds. The main analytical tool used for this purpose was an ultrasonic C-scan image analysis system to reveal this <span class="hlt">band</span>, and the defects detected in this <span class="hlt">band</span> using the preceding technique were further measured and analyzed in an optical microscope and a scanning electron microscope (SEM) using <span class="hlt">energy</span>-dispersive system (EDS). The investigation results revealed the presence of macrolevel globular macroinclusions in the size range 50 to 711 μm. The main constituents of the globular inclusions were found to be SiO2 and MnO. Impact test results revealed reduction in ductility in these <span class="hlt">bands</span> when compared with the billet material without inclusion <span class="hlt">band</span> toward the opposite face from the inner radius face. In Charpy testing, reduction in the impact <span class="hlt">energy</span> in these <span class="hlt">bands</span> was found to be 0.2 kg m. It is important to monitor the severity of macroinclusions present in these <span class="hlt">bands</span> to avoid the cracking of TMT rebars during hot rolling.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28117668','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28117668"><span><span class="hlt">Band</span> gap scaling laws in group IV nanotubes.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wang, Chongze; Fu, Xiaonan; Guo, Yangyang; Guo, Zhengxiao; Xia, Congxin; Jia, Yu</p> <p>2017-03-17</p> <p>By using the first-principles calculations, the <span class="hlt">band</span> gap properties of nanotubes formed by group IV elements have been investigated systemically. Our results reveal that for armchair nanotubes, the <span class="hlt">energy</span> gaps at K points in the Brillouin zone decrease as 1/r scaling law with the radii (r) increasing, while they are scaled by -1/r 2  + C at Γ points, here, C is a constant. Further studies show that such scaling law of K points is independent of both the chiral vector and the type of elements. Therefore, the <span class="hlt">band</span> gaps of nanotubes for a given radius can be determined by these scaling laws easily. Interestingly, we also predict the existence of indirect <span class="hlt">band</span> gap for both germanium and tin nanotubes. Our new findings provide an efficient way to determine the <span class="hlt">band</span> gaps of group IV element nanotubes by knowing the radii, as well as to facilitate the design of functional nanodevices.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29265821','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29265821"><span>A Unifying Perspective on Oxygen Vacancies in Wide <span class="hlt">Band</span> Gap Oxides.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Linderälv, Christopher; Lindman, Anders; Erhart, Paul</p> <p>2018-01-04</p> <p>Wide <span class="hlt">band</span> gap oxides are versatile materials with numerous applications in research and technology. Many properties of these materials are intimately related to defects, with the most important defect being the oxygen vacancy. Here, using electronic structure calculations, we show that the charge transition level (CTL) and eigenstates associated with oxygen vacancies, which to a large extent determine their electronic properties, are confined to a rather narrow <span class="hlt">energy</span> range, even while <span class="hlt">band</span> gap and the electronic structure of the conduction <span class="hlt">band</span> vary substantially. Vacancies are classified according to their character (deep versus shallow), which shows that the alignment of electronic eigenenergies and CTL can be understood in terms of the transition between cavity-like localized levels in the large <span class="hlt">band</span> gap limit and strong coupling between conduction <span class="hlt">band</span> and vacancy states for small to medium <span class="hlt">band</span> gaps. We consider both conventional and hybrid functionals and demonstrate that the former yields results in very good agreement with the latter provided that <span class="hlt">band</span> edge alignment is taken into account.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1981mimi.proc..330A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1981mimi.proc..330A"><span>A doubly curved reflector X-<span class="hlt">band</span> antenna with integrated IFF array</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Alia, F.; Barbati, S.</p> <p></p> <p>Primary radar antennas and Identification Friend or Foe (IFF) antennas must rotate with the same speed and synchronism, so that the target echo and IFF transponder mark will appear to the operator at the same time and at the same angular direction. A doubly-curved reflector antenna with a six-element microstrip array integrated in the reflector surface is presented to meet this requirement. The main antenna operates at X-<span class="hlt">band</span> for low angle search radar, while the secondary antenna operates at L-<span class="hlt">band</span> for IFF functions. The new configuration minimizes masking of the X-<span class="hlt">band</span> radiated <span class="hlt">energy</span> as a result of the IFF L-<span class="hlt">band</span> elements. In fact, the only effect of the microstrip array on the X-<span class="hlt">band</span> radiation pattern is the presence of several sidelobes in the + or - 90 deg angular region. The proposed new solution is compared to three other L-<span class="hlt">band/X-band</span> integrated antenna configurations, and is found to be more advantageous with respect to masking, mechanical aspects, and production costs.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010ApJ...718..920C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010ApJ...718..920C"><span>Near-infrared Thermal Emission from TrES-3b: A Ks-<span class="hlt">band</span> Detection and an H-<span class="hlt">band</span> Upper Limit on the Depth of the Secondary Eclipse</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Croll, Bryce; Jayawardhana, Ray; Fortney, Jonathan J.; Lafrenière, David; Albert, Loic</p> <p>2010-08-01</p> <p>We present H- and Ks-<span class="hlt">band</span> photometry bracketing the secondary eclipse of the hot Jupiter TrES-3b using the Wide-field Infrared Camera on the Canada-France-Hawaii Telescope. We detect the secondary eclipse of TrES-3b with a depth of 0.133+0.018 -0.016% in the Ks <span class="hlt">band</span> (8σ)—a result that is in sharp contrast to the eclipse depth reported by de Mooij & Snellen. We do not detect its thermal emission in the H <span class="hlt">band</span>, but place a 3σ limit of 0.051% on the depth of the secondary eclipse in this <span class="hlt">band</span>. A secondary eclipse of this depth in Ks requires very efficient day-to-nightside redistribution of heat and nearly isotropic reradiation, a conclusion that is in agreement with longer wavelength, mid-infrared Spitzer observations. Our 3σ upper limit on the depth of our H-<span class="hlt">band</span> secondary eclipse also argues for very efficient redistribution of heat and suggests that the atmospheric layer probed by these observations may be well homogenized. However, our H-<span class="hlt">band</span> upper limit is so constraining that it suggests the possibility of a temperature inversion at depth, or an absorbing molecule, such as methane, that further depresses the emitted flux at this wavelength. The combination of our near-infrared measurements and those obtained with Spitzer suggests that TrES-3b displays a near-isothermal dayside atmospheric temperature structure, whose spectrum is well approximated by a blackbody. We emphasize that our strict H-<span class="hlt">band</span> limit is in stark disagreement with the best-fit atmospheric model that results from longer wavelength observations only, thus highlighting the importance of near-infrared observations at multiple wavelengths, in <span class="hlt">addition</span> to those returned by Spitzer in the mid-infrared, to facilitate a comprehensive understanding of the <span class="hlt">energy</span> budgets of transiting exoplanets. Based on observations obtained with WIRCam, a joint project of CFHT, Taiwan, Korea, Canada, France, at the Canada-France-Hawaii Telescope (CFHT) which is operated by the National Research Council (NRC) of</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1408425','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1408425"><span>Multicolor emission from intermediate <span class="hlt">band</span> semiconductor ZnO 1-xSe x</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Welna, M.; Baranowski, M.; Linhart, W. M.</p> <p></p> <p>Photoluminescence and photomodulated reflectivity measurements of ZnOSe alloys are used to demonstrate a splitting of the valence <span class="hlt">band</span> due to the <span class="hlt">band</span> anticrossing interaction between localized Se states and the extended valence <span class="hlt">band</span> states of the host ZnO matrix. A strong multiband emission associated with optical transitions from the conduction <span class="hlt">band</span> to lower E - and upper E + valence subbands has been observed at room temperature. The composition dependence of the optical transition <span class="hlt">energies</span> is well explained by the electronic <span class="hlt">band</span> structure calculated using the kp method combined with the <span class="hlt">band</span> anticrossing model. The observation of the multiband emissionmore » is possible because of relatively long recombination lifetimes. Longer than 1 ns lifetimes for holes photoexcited to the lower valence subband offer a potential of using the alloy as an intermediate <span class="hlt">band</span> semiconductor for solar power conversion applications.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1408425-multicolor-emission-from-intermediate-band-semiconductor-zno1-xsex','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1408425-multicolor-emission-from-intermediate-band-semiconductor-zno1-xsex"><span>Multicolor emission from intermediate <span class="hlt">band</span> semiconductor ZnO 1-xSe x</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Welna, M.; Baranowski, M.; Linhart, W. M.; ...</p> <p>2017-03-13</p> <p>Photoluminescence and photomodulated reflectivity measurements of ZnOSe alloys are used to demonstrate a splitting of the valence <span class="hlt">band</span> due to the <span class="hlt">band</span> anticrossing interaction between localized Se states and the extended valence <span class="hlt">band</span> states of the host ZnO matrix. A strong multiband emission associated with optical transitions from the conduction <span class="hlt">band</span> to lower E - and upper E + valence subbands has been observed at room temperature. The composition dependence of the optical transition <span class="hlt">energies</span> is well explained by the electronic <span class="hlt">band</span> structure calculated using the kp method combined with the <span class="hlt">band</span> anticrossing model. The observation of the multiband emissionmore » is possible because of relatively long recombination lifetimes. Longer than 1 ns lifetimes for holes photoexcited to the lower valence subband offer a potential of using the alloy as an intermediate <span class="hlt">band</span> semiconductor for solar power conversion applications.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JAP...123k5701A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JAP...123k5701A"><span>Heterostructures with diffused interfaces: Luminescent technique for ascertainment of <span class="hlt">band</span> alignment type</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Abramkin, D. S.; Gutakovskii, A. K.; Shamirzaev, T. S.</p> <p>2018-03-01</p> <p>The experimental ascertainment of <span class="hlt">band</span> alignment type for semiconductor heterostructures with diffused interfaces is discussed. A method based on the analysis of the spectral shift of photoluminescence (PL) <span class="hlt">band</span> with excitation density (Pex) that takes into account state filling and <span class="hlt">band</span> bending effects on the PL <span class="hlt">band</span> shift is developed. It is shown that the shift of PL <span class="hlt">band</span> maximum position is proportional to ℏωmax ˜ (Ue + Uh).ln(Pex) + b.Pex1/3, where Ue (Uh) are electron (hole) Urbach <span class="hlt">energy</span> tail, and parameter b characterizes the effect of <span class="hlt">band</span> bending or is equal to zero for heterostructures with type-II or type-I <span class="hlt">band</span> alignment, respectively. The method was approved with InAs/AlAs, GaAs/AlAs, GaSb/AlAs, and AlSb/AlAs heterostructures containing quantum wells.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22573976-first-principles-energy-band-calculation-ruddlesdenpopper-compound-sr-sub-sn-sub-sub-using-modified-beckejohnson-exchange-potential','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22573976-first-principles-energy-band-calculation-ruddlesdenpopper-compound-sr-sub-sn-sub-sub-using-modified-beckejohnson-exchange-potential"><span>First-principles <span class="hlt">energy</span> <span class="hlt">band</span> calculation of Ruddlesden–Popper compound Sr{sub 3}Sn{sub 2}O{sub 7} using modified Becke–Johnson exchange potential</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Kamimura, Sunao, E-mail: kamimura-sunao@che.kyutech.ac.jp; National Institute of Advanced Industrial Science and Technology; Department of Molecular and Material Sciences, Interdisciplinary Graduate School of Engineering Science, Kyushu University, 6-1 Kasuga Kouen, Kasuga, Fukuoka 816-8580 Japan</p> <p></p> <p>The electronic structure of Sr{sub 3}Sn{sub 2}O{sub 7} is evaluated by the scalar-relativistic full potential linearized augmented plane wave (FLAPW+lo) method using the modified Becke–Johnson potential (Tran–Blaha potential) combined with the local density approximation correlation (MBJ–LDA). The fundamental gap between the valence <span class="hlt">band</span> (VB) and conduction <span class="hlt">band</span> (CB) is estimated to be 3.96 eV, which is close to the experimental value. Sn 5s states and Sr 4d states are predominant in the lower and upper CB, respectively. On the other hand, the lower VB is mainly composed of Sn 5s, 5p, and O 2p states, while the upper VB mainlymore » consists of O 2p states. These features of the DOS are well reflected by the optical transition between the upper VB and lower CB, as seen in the <span class="hlt">energy</span> dependence of the dielectric function. Furthermore, the absorption coefficient estimated from the MBJ–LDA is similar to the experimental result. - Graphical abstract: Calculated <span class="hlt">energy</span> <span class="hlt">band</span> structure along the symmetry lines of the first BZ of Sr{sub 3}Sn{sub 2}O{sub 7} crystal obtained using the MBJ potential. - Highlights: • Electronic structure of Sr{sub 3}Sn{sub 2}O{sub 7} is calculated on the basis of MBJ–LDA method for the first time. • <span class="hlt">Band</span> gap of Sr{sub 3}Sn{sub 2}O{sub 7} is determined accurately on the basis of MBJ–LDA method. • The experimental absorption spectrum of Sr{sub 3}Sn{sub 2}O{sub 7} produced by MBJ–LDA is more accurate than that obtained by GGA method.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23389265','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23389265"><span>Interface-engineering <span class="hlt">additives</span> of poly(oxyethylene tridecyl ether) for low-<span class="hlt">band</span> gap polymer solar cells consisting of PCDTBT:PCBM₇₀ bulk-heterojunction layers.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Huh, Yoon Ho; Park, Byoungchoo</p> <p>2013-01-14</p> <p>We herein report on the improved photovoltaic (PV) effects of using a polymer bulk-heterojunction (BHJ) layer that consists of a low-<span class="hlt">band</span> gap electron donor polymer of poly(N-9'-heptadecanyl-2,7-carbazole-alt-5,5-(4',7'-di-2-thienyl-2',1',3'-benzothiadiazole)) (PCDTBT) and an acceptor of [6,6]-phenyl C₇₁ butyric acid methyl ester (PCBM₇₀), doped with an interface-engineering surfactant <span class="hlt">additive</span> of poly(oxyethylene tridecyl ether) (PTE). The presence of an interface-engineering <span class="hlt">additive</span> in the PV layer results in excellent performance; the <span class="hlt">addition</span> of PTE to a PCDTBT:PCBM₇₀ system produces a power conversion efficiency (PCE) of 6.0%, which is much higher than that of a reference device without the <span class="hlt">additive</span> (4.9%). We attribute this improvement to an increased charge carrier lifetime, which is likely to be the result of the presence of PTE molecules oriented at the interfaces between the BHJ PV layer and the anode and cathode, as well as at the interfaces between the phase-separated BHJ domains. Our results suggest that the incorporation of the PTE interface-engineering <span class="hlt">additive</span> in the PCDTBT:PCBM₇₀ PV layer results in a functional composite system that shows considerable promise for use in efficient polymer BHJ PV cells.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/957418','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/957418"><span>Enhanced tunable narrow-<span class="hlt">band</span> THz emission from laser-modulated electron beams</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Xiang, D.; Stupakov, G.; /SLAC</p> <p>2009-06-19</p> <p>We propose and analyze a scheme to generate enhanced narrow-<span class="hlt">band</span> terahertz (THz) radiation through down-conversion of the frequency of optical lasers using laser-modulated electron beams. In the scheme the electron beam is first <span class="hlt">energy</span> modulated by two lasers with wave numbers k{sub 1} and k2, respectively. After passing through a dispersion section, the <span class="hlt">energy</span> modulation is converted to density modulation. Due to the nonlinear conversion process, the beam will have density modulation at wave number k = nk{sub 1} + mk{sub 2}, where n and m are positive or negative integers. By properly choosing the parameters for the lasers andmore » dispersion section, one can generate density modulation at THz frequency in the beam using optical lasers. This density-modulated beam can be used to generate powerful narrow-<span class="hlt">band</span> THz radiation. Since the THz radiation is in tight synchronization with the lasers, it should provide a high temporal resolution for the optical-pump THz-probe experiments. The central frequency of the THz radiation can be easily tuned by varying the wavelength of the two lasers and the <span class="hlt">energy</span> chirp of the electron beam. The proposed scheme is in principle able to generate intense narrow-<span class="hlt">band</span> THz radiation covering the whole THz range and offers a promising way towards the tunable intense narrow-<span class="hlt">band</span> THz sources.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5579558','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5579558"><span>Dual-<span class="hlt">Band</span> <span class="hlt">Band</span>-Pass Filter with Fixed Low <span class="hlt">Band</span> and Fluidically-Tunable High <span class="hlt">Band</span></span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Park, Eiyong; Lim, Daecheon</p> <p>2017-01-01</p> <p>In this work, we present a dual-<span class="hlt">band</span> <span class="hlt">band</span>-pass filter with fixed low-<span class="hlt">band</span> resonant frequency and tunable high-<span class="hlt">band</span> resonant frequency. The proposed filter consists of two split-ring resonators (SRRs) with a stub and microfluidic channels. The lower resonant frequency is determined by the length of the SRR alone, whereas the higher resonant frequency is determined by the lengths of the SRR and the stub. Using this characteristic, we fix the lower resonant frequency by fixing the SRR length and tune the higher resonant frequency by controlling the stub length by injecting liquid metal in the microfluidic channel. We fabricated the filter on a Duroid substrate. The microfluidic channel was made from polydimethylsiloxane (PDMS), and eutectic gallium–indium (EGaIn) was used as the liquid metal. This filter operates in two states—with, and without, the liquid metal. In the state without the liquid metal, the filter has resonant frequencies at 1.85 GHz and 3.06 GHz, with fractional bandwidths of 4.34% and 2.94%, respectively; and in the state with the liquid metal, it has resonant frequencies at 1.86 GHz and 2.98 GHz, with fractional bandwidths of 4.3% and 2.95%, respectively. PMID:28813001</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1225697','DOE-PATENT-XML'); return false;" href="https://www.osti.gov/servlets/purl/1225697"><span>Methods and <span class="hlt">energy</span> storage devices utilizing electrolytes having surface-smoothing <span class="hlt">additives</span></span></a></p> <p><a target="_blank" href="http://www.osti.gov/doepatents">DOEpatents</a></p> <p>Xu, Wu; Zhang, Jiguang; Graff, Gordon L; Chen, Xilin; Ding, Fei</p> <p>2015-11-12</p> <p>Electrodeposition and <span class="hlt">energy</span> storage devices utilizing an electrolyte having a surface-smoothing <span class="hlt">additive</span> can result in self-healing, instead of self-amplification, of initial protuberant tips that give rise to roughness and/or dendrite formation on the substrate and anode surface. For electrodeposition of a first metal (M1) on a substrate or anode from one or more cations of M1 in an electrolyte solution, the electrolyte solution is characterized by a surface-smoothing <span class="hlt">additive</span> containing cations of a second metal (M2), wherein cations of M2 have an effective electrochemical reduction potential in the solution lower than that of the cations of M1.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li class="active"><span>19</span></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_19 --> <div id="page_20" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li class="active"><span>20</span></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="381"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26647772','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26647772"><span>Broad self-trapped and slow light <span class="hlt">bands</span> based on negative refraction and interference of magnetic coupled modes.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Fang, Yun-Tuan; Ni, Zhi-Yao; Zhu, Na; Zhou, Jun</p> <p>2016-01-13</p> <p>We propose a new mechanism to achieve light localization and slow light. Through the study on the coupling of two magnetic surface modes, we find a special convex <span class="hlt">band</span> that takes on a negative refraction effect. The negative refraction results in an <span class="hlt">energy</span> flow concellation effect from two degenerated modes on the convex <span class="hlt">band</span>. The <span class="hlt">energy</span> flow concellation effect leads to forming of the self-trapped and slow light <span class="hlt">bands</span>. In the self-trapped <span class="hlt">band</span> light is localized around the source without reflection wall in the waveguide direction, whereas in the slow light <span class="hlt">band</span>, light becomes the standing-waves and moving standing-waves at the center and the two sides of the waveguide, respectively.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20110020799','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20110020799"><span>Interpreting Methanol v(sub 2)-<span class="hlt">Band</span> Emission in Comets Using Empirical Fluorescence g-Factors</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>DiSanti, Michael; Villanueva, G. L.; Bonev, B. P.; Mumma, M. J.; Paganini, L.; Gibb, E. L.; Magee-Sauer, K.</p> <p>2011-01-01</p> <p>For many years we have been developing the ability, through high-resolution spectroscopy targeting ro-vibrational emission in the approximately 3 - 5 micrometer region, to quantify a suite of (approximately 10) parent volatiles in comets using quantum mechanical fluorescence models. Our efforts are ongoing and our latest includes methanol (CH3OH). This is unique among traditionally targeted species in having lacked sufficiently robust models for its symmetric (v(sub 3) <span class="hlt">band</span>) and asymmetric (v(sub 2) and v(sub 9) <span class="hlt">bands</span>) C-H3 stretching modes, required to provide accurate predicted intensities for individual spectral lines and hence rotational temperatures and production rates. This has provided the driver for undertaking a detailed empirical study of line intensities, and has led to substantial progress regarding our ability to interpret CH3OH in comets. The present study concentrates on the spectral region from approximately 2970 - 3010 per centimeter (3.367 - 3.322 micrometer), which is dominated by emission in the (v(sub 7) <span class="hlt">band</span> of C2H6 and the v(sub 2) <span class="hlt">band</span> of CH3OH, with minor contributions from CH3OH (v(sub 9) <span class="hlt">band</span>), CH4 (v(sub 3)), and OH prompt emissions (v(sub 1) and v(sub 2)- v(sub 1)). Based on laboratory jet-cooled spectra (at a rotational temperature near 20 K)[1], we incorporated approximately 100 lines of the CH3OH v(sub 2) <span class="hlt">band</span>, having known frequencies and lower state rotational <span class="hlt">energies</span>, into our model. Line intensities were determined through comparison with several comets we observed with NIRSPEC at Keck 2, after removal of continuum and <span class="hlt">additional</span> molecular emissions and correcting for atmospheric extinction. In <span class="hlt">addition</span> to the above spectral region, NIRSPEC allows simultaneous sampling of the CH3OH v(sub 3) <span class="hlt">band</span> (centered at 2844 per centimeter, or 3.516 micrometers and several hot <span class="hlt">bands</span> of H2O in the approximately 2.85 - 2.9 micrometer region, at a nominal spectral resolving power of approximately 25,000 [2]. Empirical g-factors for v(sub 2</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/10198957','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/10198957"><span>[Effect of biologically active food <span class="hlt">additives</span> on <span class="hlt">energy</span> metabolism and human body weight].</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Gapparov, M M</p> <p>1999-01-01</p> <p>Review is devoted to analysis of human <span class="hlt">energy</span> requirements depending on age, sex, occupational and living condition. Special attention was paid to importance of strict balance in organism between consumption and expense of <span class="hlt">energy</span>. Modern views on mechanism of action food supplements as <span class="hlt">additional</span> instrument of regulation of <span class="hlt">energy</span> metabolism for correction of surplus body weight is given. Review is the first attempt of systematisation of biologically active food supplements according to their mechanism of action both on nutrition processes and on biochemical mechanisms of assimilation and utilisation of macronutrients, in particular of fats and carbohydrates.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23396813','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23396813"><span>Quasiparticle semiconductor <span class="hlt">band</span> structures including spin-orbit interactions.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Malone, Brad D; Cohen, Marvin L</p> <p>2013-03-13</p> <p>We present first-principles calculations of the quasiparticle <span class="hlt">band</span> structure of the group IV materials Si and Ge and the group III-V compound semiconductors AlP, AlAs, AlSb, InP, InAs, InSb, GaP, GaAs and GaSb. Calculations are performed using the plane wave pseudopotential method and the 'one-shot' GW method, i.e. G(0)W(0). Quasiparticle <span class="hlt">band</span> structures, augmented with the effects of spin-orbit, are obtained via a Wannier interpolation of the obtained quasiparticle <span class="hlt">energies</span> and calculated spin-orbit matrix. Our calculations explicitly treat the shallow semicore states of In and Ga, which are known to be important in the description of the electronic properties, as valence states in the quasiparticle calculation. Our calculated quasiparticle <span class="hlt">energies</span>, combining both the ab initio evaluation of the electron self-<span class="hlt">energy</span> and the vector part of the pseudopotential representing the spin-orbit effects, are in generally very good agreement with experimental values. These calculations illustrate the predictive power of the methodology as applied to group IV and III-V semiconductors.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/873026','DOE-PATENT-XML'); return false;" href="https://www.osti.gov/servlets/purl/873026"><span>Thermophotovoltaic conversion using selective infrared line emitters and large <span class="hlt">band</span> gap photovoltaic devices</span></a></p> <p><a target="_blank" href="http://www.osti.gov/doepatents">DOEpatents</a></p> <p>Brandhorst, Jr., Henry W.; Chen, Zheng</p> <p>2000-01-01</p> <p>Efficient thermophotovoltaic conversion can be performed using photovoltaic devices with a <span class="hlt">band</span> gap in the 0.75-1.4 electron volt range, and selective infrared emitters chosen from among the rare earth oxides which are thermally stimulated to emit infrared radiation whose <span class="hlt">energy</span> very largely corresponds to the aforementioned <span class="hlt">band</span> gap. It is possible to use thermovoltaic devices operating at relatively high temperatures, up to about 300.degree. C., without seriously impairing the efficiency of <span class="hlt">energy</span> conversion.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22493003-ultra-wide-acoustic-band-gaps-pillar-based-phononic-crystal-strips','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22493003-ultra-wide-acoustic-band-gaps-pillar-based-phononic-crystal-strips"><span>Ultra-wide acoustic <span class="hlt">band</span> gaps in pillar-based phononic crystal strips</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Coffy, Etienne, E-mail: etienne.coffy@femto-st.fr; Lavergne, Thomas; Addouche, Mahmoud</p> <p>2015-12-07</p> <p>An original approach for designing a one dimensional phononic crystal strip with an ultra-wide <span class="hlt">band</span> gap is presented. The strip consists of periodic pillars erected on a tailored beam, enabling the generation of a <span class="hlt">band</span> gap that is due to both Bragg scattering and local resonances. The optimized combination of both effects results in the lowering and the widening of the main <span class="hlt">band</span> gap, ultimately leading to a gap-to-midgap ratio of 138%. The design method used to improve the <span class="hlt">band</span> gap width is based on the flattening of phononic <span class="hlt">bands</span> and relies on the study of the modal <span class="hlt">energy</span> distributionmore » within the unit cell. The computed transmission through a finite number of periods corroborates the dispersion diagram. The strong attenuation, in excess of 150 dB for only five periods, highlights the interest of such ultra-wide <span class="hlt">band</span> gap phononic crystal strips.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AnPhy.382..160Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AnPhy.382..160Z"><span>New edge-centered photonic square lattices with flat <span class="hlt">bands</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zhang, Da; Zhang, Yiqi; Zhong, Hua; Li, Changbiao; Zhang, Zhaoyang; Zhang, Yanpeng; Belić, Milivoj R.</p> <p>2017-07-01</p> <p>We report a new class of edge-centered photonic square lattices with multiple flat <span class="hlt">bands</span>, and consider in detail two examples: the Lieb-5 and Lieb-7 lattices. In these lattices, there are 5 and 7 sites in the unit cell and in general, the number is restricted to odd integers. The number of flat <span class="hlt">bands</span> m in the new Lieb lattices is related to the number of sites N in the unit cell by a simple formula m =(N - 1) / 2. The flat <span class="hlt">bands</span> reported here are independent of the pseudomagnetic field. The properties of lattices with even and odd number of flat <span class="hlt">bands</span> are different. We consider the localization of light in such Lieb lattices. If the input beam excites the flat-<span class="hlt">band</span> mode, it will not diffract during propagation, owing to the strong mode localization. In the Lieb-7 lattice, the beam will also oscillate during propagation and still not diffract. The period of oscillation is determined by the <span class="hlt">energy</span> difference between the two flat <span class="hlt">bands</span>. This study provides a new platform for investigating light trapping, photonic topological insulators, and pseudospin-mediated vortex generation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1356148-strain-fields-induced-kink-band-propagation-cu-nb-nanolaminate-composites','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1356148-strain-fields-induced-kink-band-propagation-cu-nb-nanolaminate-composites"><span>Strain fields induced by kink <span class="hlt">band</span> propagation in Cu-Nb nanolaminate composites</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Nizolek, T. J.; Begley, M. R.; McCabe, R. J.; ...</p> <p>2017-07-01</p> <p>Kink <span class="hlt">band</span> formation is a common deformation mode for anisotropic materials and has been observed in polymer matrix fiber composites, single crystals, geological formations, and recently in metallic nanolaminates. While numerous studies have been devoted to kink <span class="hlt">band</span> formation, the majority do not consider the often rapid and unstable process of kink <span class="hlt">band</span> propagation. In this paper, we take advantage of stable kink <span class="hlt">band</span> formation in Cu-Nb nanolaminates to quantitatively map the local strain fields surrounding a propagating kink <span class="hlt">band</span> during uniaxial compression. Kink <span class="hlt">bands</span> are observed to initiate at specimen edges, propagate across the sample during a rising globalmore » stress, and induce extended strain fields in the non-kinked material surrounding the propagating kink <span class="hlt">band</span>. Finally, it is proposed that these stress/strain fields significantly contribute to the total <span class="hlt">energy</span> dissipated during kinking and, analogous to crack tip stress/strain fields, influence the direction of kink propagation and therefore the kink <span class="hlt">band</span> inclination angle.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017OptCo.382..151K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017OptCo.382..151K"><span>Triple-<span class="hlt">band</span> metamaterial absorption utilizing single rectangular hole</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kim, Seung Jik; Yoo, Young Joon; Kim, Young Ju; Lee, YoungPak</p> <p>2017-01-01</p> <p>In the general metamaterial absorber, the single absorption <span class="hlt">band</span> is made by the single meta-pattern. Here, we introduce the triple-<span class="hlt">band</span> metamaterial absorber only utilizing single rectangular hole. We also demonstrate the absorption mechanism of the triple absorption. The first absorption peak was caused by the fundamental magnetic resonance in the metallic part between rectangular holes. The second absorption was generated by induced tornado magnetic field. The process of realizing the second <span class="hlt">band</span> is also presented. The third absorption was induced by the third-harmonic magnetic resonance in the metallic region between rectangular holes. In <span class="hlt">addition</span>, the visible-range triple-<span class="hlt">band</span> absorber was also realized by using similar but smaller single rectangular-hole structure. These results render the simple metamaterials for high frequency in large scale, which can be useful in the fabrication of metamaterials operating in the optical range.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018ApPhA.124..364Q','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018ApPhA.124..364Q"><span><span class="hlt">Band</span> gap structures for 2D phononic crystals with composite scatterer</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Qi, Xiao-qiao; Li, Tuan-jie; Zhang, Jia-long; Zhang, Zhen; Tang, Ya-qiong</p> <p>2018-05-01</p> <p>We investigated the <span class="hlt">band</span> gap structures in two-dimensional phononic crystals with composite scatterer. The composite scatterers are composed of two materials (Bragg scattering type) or three materials (locally resonance type). The finite element method is used to calculate the <span class="hlt">band</span> gap structure, eigenmodes and transmission spectrum. The variation of the location and width of <span class="hlt">band</span> gap are also investigated as a function of material ratio in the scatterer. We have found that the change trends the widest <span class="hlt">band</span> gap of the two phononic crystals are different as the material ratio changing. In <span class="hlt">addition</span> to this, there are three complete <span class="hlt">band</span> gaps at most for the Bragg-scattering-type phononic crystals in the first six <span class="hlt">bands</span>; however, the locally resonance-type phononic crystals exist only two complete <span class="hlt">band</span> gap at most in the first six <span class="hlt">bands</span>. The gap-tuning effect can be controlled by the material ratio in the scatterer.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018PhRvB..97l5121M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PhRvB..97l5121M"><span>Interplay of Coulomb interactions and disorder in three-dimensional quadratic <span class="hlt">band</span> crossings without time-reversal symmetry and with unequal masses for conduction and valence <span class="hlt">bands</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mandal, Ipsita; Nandkishore, Rahul M.</p> <p>2018-03-01</p> <p>Coulomb interactions famously drive three-dimensional quadratic <span class="hlt">band</span> crossing semimetals into a non-Fermi liquid phase of matter. In a previous work [Nandkishore and Parameswaran, Phys. Rev. B 95, 205106 (2017), 10.1103/PhysRevB.95.205106], the effect of disorder on this non-Fermi liquid phase was investigated, assuming that the <span class="hlt">band</span> structure was isotropic, assuming that the conduction and valence <span class="hlt">bands</span> had the same <span class="hlt">band</span> mass, and assuming that the disorder preserved exact time-reversal symmetry and statistical isotropy. It was shown that the non-Fermi liquid fixed point is unstable to disorder and that a runaway flow to strong disorder occurs. In this paper, we extend that analysis by relaxing the assumption of time-reversal symmetry and allowing the electron and hole masses to differ (but continuing to assume isotropy of the low <span class="hlt">energy</span> <span class="hlt">band</span> structure). We first incorporate time-reversal symmetry breaking disorder and demonstrate that there do not appear any new fixed points. Moreover, while the system continues to flow to strong disorder, time-reversal-symmetry-breaking disorder grows asymptotically more slowly than time-reversal-symmetry-preserving disorder, which we therefore expect should dominate the strong-coupling phase. We then allow for unequal electron and hole masses. We show that whereas asymmetry in the two masses is irrelevant in the clean system, it is relevant in the presence of disorder, such that the `effective masses' of the conduction and valence <span class="hlt">bands</span> should become sharply distinct in the low-<span class="hlt">energy</span> limit. We calculate the RG flow equations for the disordered interacting system with unequal <span class="hlt">band</span> masses and demonstrate that the problem exhibits a runaway flow to strong disorder. Along the runaway flow, time-reversal-symmetry-preserving disorder grows asymptotically more rapidly than both time-reversal-symmetry-breaking disorder and the Coulomb interaction.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22390743-band-bending-ferroelectric-surfaces-interfaces-investigated-ray-photoelectron-spectroscopy','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22390743-band-bending-ferroelectric-surfaces-interfaces-investigated-ray-photoelectron-spectroscopy"><span><span class="hlt">Band</span> bending at ferroelectric surfaces and interfaces investigated by x-ray photoelectron spectroscopy</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Apostol, Nicoleta Georgiana, E-mail: nicoleta.apostol@infim.ro</p> <p>2014-11-24</p> <p>This work reports on the use of X-ray photoelectron spectroscopy to quantify <span class="hlt">band</span> bending at ferroelectric free surfaces and at their interfaces with metals. Surfaces exhibiting out-of-plane ferroelectric polarization are characterized by a <span class="hlt">band</span> bending, due to the formation of a dipole layer at the surface, composed by the uncompensated polarization charges (due to ionic displacement) and to the depolarization charge sheet of opposite sign, composed by mobile charge carriers, which migrate near surface, owing to the depolarization electric field. To this surface <span class="hlt">band</span> bending due to out-of-plane polarization states, metal-semiconductor Schottky barriers must be considered <span class="hlt">additionally</span> when ferroelectrics aremore » covered by metal layers. It is found that the net <span class="hlt">band</span> bending is not always an algebraic sum of the two effects discussed above, since sometimes the metal is able to provide <span class="hlt">additional</span> charge carriers, which are able to fully compensate the surface charge of the ferroelectric, up to the vanishing of the ferroelectric <span class="hlt">band</span> bending. The two cases which will be discussed in more detail are Au and Cu deposited by molecular beam epitaxy on PbZr{sub 0.2}Ti{sub 0.8}O{sub 3}(001) single crystal thin layers, prepared by pulsed laser deposition. Gold forms unconnected nanoparticles, and their effect on the <span class="hlt">band</span> bending is the apparition of a Schottky <span class="hlt">band</span> bending <span class="hlt">additional</span> to the <span class="hlt">band</span> bending due to the out-of-plane polarization. Copper, starting with a given thickness, forms continuous metal layers connected to the ground of the system, and provide electrons in sufficient quantity to compensate the <span class="hlt">band</span> bending due to the out-of-plane polarization.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4213712','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4213712"><span>The Influence of Feed <span class="hlt">Energy</span> Density and a Formulated <span class="hlt">Additive</span> on Rumen and Rectal Temperature in Hanwoo Steers</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Cho, Sangbuem; Mbiriri, David Tinotenda; Shim, Kwanseob; Lee, A-Leum; Oh, Seong-Jin; Yang, Jinho; Ryu, Chaehwa; Kim, Young-Hoon; Seo, Kang-Seok; Chae, Jung-Il; Oh, Young Kyoon; Choi, Nag-Jin</p> <p>2014-01-01</p> <p>The present study investigated the optimum blending condition of protected fat, choline and yeast culture for lowering of rumen temperature. The Box Benken experimental design, a fractional factorial arrangement, and response surface methodology were employed. The optimum blending condition was determined using the rumen simulated in vitro fermentation. An <span class="hlt">additive</span> formulated on the optimum condition contained 50% of protected fat, 25% of yeast culture, 5% of choline, 7% of organic zinc, 6.5% of cinnamon, and 6.5% of stevioside. The feed <span class="hlt">additive</span> was supplemented at a rate of 0.1% of diet (orchard grass:concentrate, 3:7) and compared with a control which had no <span class="hlt">additive</span>. The treatment resulted in lower volatile fatty acid (VFA) concentration and biogas than the control. To investigate the effect of the optimized <span class="hlt">additive</span> and feed <span class="hlt">energy</span> levels on rumen and rectal temperatures, four rumen cannulated Hanwoo (Korean native beef breed) steers were in a 4×4 Latin square design. <span class="hlt">Energy</span> levels were varied to low and high by altering the ratio of forage to concentrate in diet: low <span class="hlt">energy</span> (6:4) and high <span class="hlt">energy</span> (4:6). The <span class="hlt">additive</span> was added at a rate of 0.1% of the diet. The following parameters were measured; feed intake, rumen and rectal temperatures, ruminal pH and VFA concentration. This study was conducted in an environmentally controlled house with temperature set at 30°C and relative humidity levels of 70%. Steers were housed individually in raised crates to facilitate collection of urine and feces. The adaptation period was for 14 days, 2 days for sampling and 7 days for resting the animals. The <span class="hlt">additive</span> significantly reduced both rumen (p<0.01) and rectal temperatures (p<0.001) without depressed feed intake. There were interactions (p<0.01) between <span class="hlt">energy</span> level and <span class="hlt">additive</span> on ruminal temperature. Neither <span class="hlt">additive</span> nor <span class="hlt">energy</span> level had an effect on total VFA concentration. The <span class="hlt">additive</span> however, significantly increased (p<0.01) propionate and subsequently had lower</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25358327','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25358327"><span>The influence of feed <span class="hlt">energy</span> density and a formulated <span class="hlt">additive</span> on rumen and rectal temperature in hanwoo steers.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Cho, Sangbuem; Mbiriri, David Tinotenda; Shim, Kwanseob; Lee, A-Leum; Oh, Seong-Jin; Yang, Jinho; Ryu, Chaehwa; Kim, Young-Hoon; Seo, Kang-Seok; Chae, Jung-Il; Oh, Young Kyoon; Choi, Nag-Jin</p> <p>2014-11-01</p> <p>The present study investigated the optimum blending condition of protected fat, choline and yeast culture for lowering of rumen temperature. The Box Benken experimental design, a fractional factorial arrangement, and response surface methodology were employed. The optimum blending condition was determined using the rumen simulated in vitro fermentation. An <span class="hlt">additive</span> formulated on the optimum condition contained 50% of protected fat, 25% of yeast culture, 5% of choline, 7% of organic zinc, 6.5% of cinnamon, and 6.5% of stevioside. The feed <span class="hlt">additive</span> was supplemented at a rate of 0.1% of diet (orchard grass:concentrate, 3:7) and compared with a control which had no <span class="hlt">additive</span>. The treatment resulted in lower volatile fatty acid (VFA) concentration and biogas than the control. To investigate the effect of the optimized <span class="hlt">additive</span> and feed <span class="hlt">energy</span> levels on rumen and rectal temperatures, four rumen cannulated Hanwoo (Korean native beef breed) steers were in a 4×4 Latin square design. <span class="hlt">Energy</span> levels were varied to low and high by altering the ratio of forage to concentrate in diet: low <span class="hlt">energy</span> (6:4) and high <span class="hlt">energy</span> (4:6). The <span class="hlt">additive</span> was added at a rate of 0.1% of the diet. The following parameters were measured; feed intake, rumen and rectal temperatures, ruminal pH and VFA concentration. This study was conducted in an environmentally controlled house with temperature set at 30°C and relative humidity levels of 70%. Steers were housed individually in raised crates to facilitate collection of urine and feces. The adaptation period was for 14 days, 2 days for sampling and 7 days for resting the animals. The <span class="hlt">additive</span> significantly reduced both rumen (p<0.01) and rectal temperatures (p<0.001) without depressed feed intake. There were interactions (p<0.01) between <span class="hlt">energy</span> level and <span class="hlt">additive</span> on ruminal temperature. Neither <span class="hlt">additive</span> nor <span class="hlt">energy</span> level had an effect on total VFA concentration. The <span class="hlt">additive</span> however, significantly increased (p<0.01) propionate and subsequently had lower</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2005cmns.conf..735C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2005cmns.conf..735C"><span>Nuts and Bolts of the Ion <span class="hlt">Band</span> State Theory</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Chubb, Scott R.</p> <p>2005-12-01</p> <p>The nuts and bolts of our ion <span class="hlt">band</span> state theory of low <span class="hlt">energy</span> nuclear reactions (LENR's) in palladium-deuteride (PdD) and palladium-hydride (PdH) are the electrons that hold together or tear apart the bonds (or lack of bonds) between deuterons (d's) or protons (p's) and the host material. In PdDx and PdHx, this bonding is strongly correlated with loading. In ambient loading conditions (x ≲ 0.6), bonding inhibits ion <span class="hlt">band</span> state occupation. As x → 1, slight increases and decreases in loading can induce "vibrations" (which have conventionally been thought to occur from phonons) that can induce potential losses or increases of p/d. Naive assumptions about phonons fail to include these losses and increases. These effects can occur because neither H or D has core electrons and because in either PdD or PdH, the electrons near the Fermi <span class="hlt">energy</span> have negligible overlap with the nucleus of either D or H. In the past, implicitly, we have used these facts to justify our ion <span class="hlt">band</span> state theory. Here, we present a more formal justification, based on the relationship between H(D) ion <span class="hlt">band</span> states (IBS's) and H(D) phonons that includes a microscopic picture that explains why occupation of IBS's can occur in PdD and PdH and how this can lead to nuclear reactions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1097780','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1097780"><span>Large-scale Manufacturing of Nanoparticulate-based Lubrication <span class="hlt">Additives</span> for Improved <span class="hlt">Energy</span> Efficiency and Reduced Emissions</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Erdemir, Ali</p> <p></p> <p>This project was funded under the Department of <span class="hlt">Energy</span> (DOE) Lab Call on Nanomanufacturing for <span class="hlt">Energy</span> Efficiency and was directed toward the development of novel boron-based nanocolloidal lubrication <span class="hlt">additives</span> for improving the friction and wear performance of machine components in a wide range of industrial and transportation applications. Argonne's research team concentrated on the scientific and technical aspects of the project, using a range of state-of-the art analytical and tribological test facilities. Argonne has extensive past experience and expertise in working with boron-based solid and liquid lubrication <span class="hlt">additives</span>, and has intellectual property ownership of several. There were two industrial collaboratorsmore » in this project: Ashland Oil (represented by its Valvoline subsidiary) and Primet Precision Materials, Inc. (a leading nanomaterials company). There was also a sub-contract with the University of Arkansas. The major objectives of the project were to develop novel boron-based nanocolloidal lubrication <span class="hlt">additives</span> and to optimize and verify their performance under boundary-lubricated sliding conditions. The project also tackled problems related to colloidal dispersion, larger-scale manufacturing and blending of nano-<span class="hlt">additives</span> with base carrier oils. Other important issues dealt with in the project were determination of the optimum size and concentration of the particles and compatibility with various base fluids and/or <span class="hlt">additives</span>. Boron-based particulate <span class="hlt">additives</span> considered in this project included boric acid (H{sub 3}BO{sub 3}), hexagonal boron nitride (h-BN), boron oxide, and borax. As part of this project, we also explored a hybrid MoS{sub 2} + boric acid formulation approach for more effective lubrication and reported the results. The major motivation behind this work was to reduce <span class="hlt">energy</span> losses related to friction and wear in a wide spectrum of mechanical systems and thereby reduce our dependence on imported oil. Growing concern over</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22590585-importance-band-tail-recombination-current-collection-open-circuit-voltage-cztsse-solar-cells','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22590585-importance-band-tail-recombination-current-collection-open-circuit-voltage-cztsse-solar-cells"><span>The importance of <span class="hlt">band</span> tail recombination on current collection and open-circuit voltage in CZTSSe solar cells</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Moore, James E.; Purdue University, West Lafayette, Indiana 47907; Hages, Charles J.</p> <p>2016-07-11</p> <p>Cu{sub 2}ZnSn(S,Se){sub 4} (CZTSSe) solar cells typically exhibit high short-circuit current density (J{sub sc}), but have reduced cell efficiencies relative to other thin film technologies due to a deficit in the open-circuit voltage (V{sub oc}), which prevent these devices from becoming commercially competitive. Recent research has attributed the low V{sub oc} in CZTSSe devices to small scale disorder that creates <span class="hlt">band</span> tail states within the absorber <span class="hlt">band</span> gap, but the physical processes responsible for this V{sub oc} reduction have not been elucidated. In this paper, we show that carrier recombination through non-mobile <span class="hlt">band</span> tail states has a strong voltage dependencemore » and is a significant performance-limiting factor, and including these effects in simulation allows us to simultaneously explain the V{sub oc} deficit, reduced fill factor, and voltage-dependent quantum efficiency with a self-consistent set of material parameters. Comparisons of numerical simulations to measured data show that reasonable values for the <span class="hlt">band</span> tail parameters (characteristic <span class="hlt">energy</span>, capture rate) can account for the observed low V{sub oc}, high J{sub sc}, and voltage dependent collection efficiency. These results provide <span class="hlt">additional</span> evidence that the presence of <span class="hlt">band</span> tail states accounts for the low efficiencies of CZTSSe solar cells and further demonstrates that recombination through non-mobile <span class="hlt">band</span> tail states is the dominant efficiency limiting mechanism.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2000JGR...10520857C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2000JGR...10520857C"><span>Nitric oxide excited under auroral conditions: Excited state densities and <span class="hlt">band</span> emissions</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Cartwright, D. C.; Brunger, M. J.; Campbell, L.; Mojarrabi, B.; Teubner, P. J. O.</p> <p>2000-09-01</p> <p>Electron impact excitation of vibrational levels in the ground electronic state and nine excited electronic states in NO has been simulated for an IBC II aurora (i.e., ˜10 kR in 3914 Å radiation) in order to predict NO excited state number densities and <span class="hlt">band</span> emission intensities. New integral electron impact excitation cross sections for NO were combined with a measured IBC II auroral secondary electron distribution, and the vibrational populations of 10 NO electronic states were determined under conditions of statistical equilibrium. This model predicts an extended vibrational distribution in the NO ground electronic state produced by radiative cascade from the seven higher-lying doublet excited electronic states populated by electron impact. In <span class="hlt">addition</span> to significant <span class="hlt">energy</span> storage in vibrational excitation of the ground electronic state, both the a 4Π and L2 Φ excited electronic states are predicted to have relatively high number densities because they are only weakly connected to lower electronic states by radiative decay. Fundamental mode radiative transitions involving the lowest nine excited vibrational levels in the ground electronic state are predicted to produce infrared (IR) radiation from 5.33 to 6.05 μm with greater intensity than any single NO electronic emission <span class="hlt">band</span>. Fundamental mode radiative transitions within the a 4Π electronic state, in the 10.08-11.37 μm region, are predicted to have IR intensities comparable to individual electronic emission <span class="hlt">bands</span> in the Heath and ɛ <span class="hlt">band</span> systems. Results from this model quantitatively predict the vibrational quantum number dependence of the NO IR measurements of Espy et al. [1988].</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29687571','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29687571"><span>The size effect to O2- -Ce4+ charge transfer emission and <span class="hlt">band</span> gap structure of Sr2 CeO4.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wang, Wenjun; Pan, Yu; Zhang, Wenying; Liu, Xiaoguang; Li, Ling</p> <p>2018-04-24</p> <p>Sr 2 CeO 4 phosphors with different crystalline sizes were synthesized by the sol-gel method or the solid-state reaction. Their crystalline size, luminescence intensity of O 2- -Ce 4+ charge transfer and <span class="hlt">energy</span> gaps were obtained through the characterization by X-ray diffraction, photoluminescence spectra, as well as UV-visible diffuse reflectance measurements. An inverse relationship between photoluminescence (PL) spectra and crystalline size was observed when the heating temperature was from 1000°C to 1300°C. In <span class="hlt">addition</span>, <span class="hlt">band</span> <span class="hlt">energy</span> calculated for all samples showed that a reaction temperature of 1200°C for the solid-state method and 1100°C for sol-gel method gave the largest values, which corresponded with the smallest crystalline size. Correlation between PL intensity and crystalline size showed an inverse relationship. <span class="hlt">Band</span> structure, density of states and partial density of states of the crystal were calculated to analyze the mechanism using the cambrige sequential total <span class="hlt">energy</span> package (CASTEP) module integrated with Materials Studio software. Copyright © 2018 John Wiley & Sons, Ltd.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22590494-band-gap-tuning-amorphous-al-oxides-zr-alloying','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22590494-band-gap-tuning-amorphous-al-oxides-zr-alloying"><span><span class="hlt">Band</span> gap tuning of amorphous Al oxides by Zr alloying</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Canulescu, S., E-mail: stec@fotonik.dtu.dk; Schou, J.; Jones, N. C.</p> <p>2016-08-29</p> <p>The optical <span class="hlt">band</span> gap and electronic structure of amorphous Al-Zr mixed oxides with Zr content ranging from 4.8 to 21.9% were determined using vacuum ultraviolet and X-ray absorption spectroscopy. The light scattering by the nano-porous structure of alumina at low wavelengths was estimated based on the Mie scattering theory. The dependence of the optical <span class="hlt">band</span> gap of the Al-Zr mixed oxides on the Zr content deviates from linearity and decreases from 7.3 eV for pure anodized Al{sub 2}O{sub 3} to 6.45 eV for Al-Zr mixed oxides with a Zr content of 21.9%. With increasing Zr content, the conduction <span class="hlt">band</span> minimum changes non-linearlymore » as well. Fitting of the <span class="hlt">energy</span> <span class="hlt">band</span> gap values resulted in a bowing parameter of ∼2 eV. The <span class="hlt">band</span> gap bowing of the mixed oxides is assigned to the presence of the Zr d-electron states localized below the conduction <span class="hlt">band</span> minimum of anodized Al{sub 2}O{sub 3}.« less</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li class="active"><span>20</span></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_20 --> <div id="page_21" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li class="active"><span>21</span></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="401"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1999JMoSp.197..188W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1999JMoSp.197..188W"><span>Identification of New Hot <span class="hlt">Bands</span> in the Blue and Green <span class="hlt">Band</span> Systems of FeH</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wilson, Catherine; Brown, John M.</p> <p>1999-10-01</p> <p>A particularly rich region of the electronic spectrum of FeH from 525 to 545 nm was investigated using the techniques of dispersed and undispersed laser-induced fluorescence. Analysis has led to the discovery that several different electronic transitions are embedded in this region; the (0, 0) and (1, 1) <span class="hlt">bands</span> of the e6Π-a6Δ (green) system, the (0, 2) <span class="hlt">band</span> of the g6Φ-X4Δ (intercombination) system, the (0, 1) <span class="hlt">band</span> of the g6Φ-a6Δ (blue) system, and the (0, 0) <span class="hlt">band</span> of the g6Φ-b6Π system. Seventy-five lines were assigned in the (0, 1) <span class="hlt">band</span> of the g6Φ-a6Δ transition. These, with the assignment of an <span class="hlt">additional</span> 14 lines in the 583 nm region to the (0, 1) <span class="hlt">band</span> of the e6Π-a6Δ transition, led to the extension of the known term values to higher J values for the Ω = 9/2, 7/2, and 5/2 spin components of the v = 1 level of the a6Δ state and the novel characterization of the a6Δ3/2 (v = 1) and g6Φ5/2 (v = 0) components. A further 73 lines were assigned to the first four subbands of the (1, 1) <span class="hlt">band</span> of the e6Π-a6Δ transition and term values for the lowest four spin components of the v = 1 level of the e6Π state were determined. This provides the first experimental measurement of a vibrational interval in one of the higher lying electronic states of FeH. The interval does not appear to vary strongly between the spin components (ΔG1/2 = 1717, 1713, 1710 cm-1 for Ω = 7/2, 5/2, 3/2, respectively). Remarkably few of the hot-<span class="hlt">band</span> transitions assigned in this work could be identified in the complex, high-temperature spectrum of FeH recorded by P. McCormack and S. O'Connor [Astron. Astrophys. Suppl. 26, 373-380 (1976)].</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017OptCo.392...31K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017OptCo.392...31K"><span>Microwave <span class="hlt">energy</span> harvesting based on metamaterial absorbers with multi-layered square split rings for wireless communications</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Karaaslan, Muharrem; Bağmancı, Mehmet; Ünal, Emin; Akgol, Oguzhan; Sabah, Cumali</p> <p>2017-06-01</p> <p>We propose the design of a multiband absorber based on multi-layered square split ring (MSSR) structure. The multi-layered metamaterial structure is designed to be used in the frequency <span class="hlt">bands</span> such as WIMAX, WLAN and satellite communication region. The absorption levels of the proposed structure are higher than 90% for all resonance frequencies. In <span class="hlt">addition</span>, the incident angle and polarization dependence of the multi-layered metamaterial absorber and harvester is also investigated and it is observed that the structure has polarization angle independent frequency response with good absorption characteristics in the entire working frequency <span class="hlt">band</span>. The <span class="hlt">energy</span> harvesting ratios of the structure is investigated especially for the resonance frequencies at which the maximum absorption occurs. The <span class="hlt">energy</span> harvesting potential of the proposed MSSRs is as good as those of the structures given in the literature. Therefore, the suggested design having good absorption, polarization and angle independent characteristics with a wide bandwidth is a potential candidate for future <span class="hlt">energy</span> harvesting applications in commonly used wireless communication <span class="hlt">bands</span>, namely WIMAX, WLAN and satellite communication <span class="hlt">bands</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016ZNatA..71..493Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016ZNatA..71..493Y"><span><span class="hlt">Band</span> Structure Characteristics of Nacreous Composite Materials with Various Defects</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yin, J.; Zhang, S.; Zhang, H. W.; Chen, B. S.</p> <p>2016-06-01</p> <p>Nacreous composite materials have excellent mechanical properties, such as high strength, high toughness, and wide phononic <span class="hlt">band</span> gap. In order to research <span class="hlt">band</span> structure characteristics of nacreous composite materials with various defects, supercell models with the Brick-and-Mortar microstructure are considered. An efficient multi-level substructure algorithm is employed to discuss the <span class="hlt">band</span> structure. Furthermore, two common systems with point and line defects and varied material parameters are discussed. In <span class="hlt">addition</span>, <span class="hlt">band</span> structures concerning straight and deflected crack defects are calculated by changing the shear modulus of the mortar. Finally, the sensitivity of <span class="hlt">band</span> structures to the random material distribution is presented by considering different volume ratios of the brick. The results reveal that the first <span class="hlt">band</span> gap of a nacreous composite material is insensitive to defects under certain conditions. It will be of great value to the design and synthesis of new nacreous composite materials for better dynamic properties.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1338137-topological-nonsymmorphic-metals-from-band-inversion','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1338137-topological-nonsymmorphic-metals-from-band-inversion"><span>Topological nonsymmorphic metals from <span class="hlt">band</span> inversion</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Muechler, Lukas; Alexandradinata, A.; Neupert, Titus; ...</p> <p>2016-12-29</p> <p>Here, we expand the phase diagram of two-dimensional, nonsymmorphic crystals at integer fillings that do not guarantee gaplessness. In <span class="hlt">addition</span> to the trivial, gapped phase that is expected, we find that <span class="hlt">band</span> inversion leads to a class of topological, gapless phases. These topological phases are exemplified by the monolayers of MTe 2 (M ¼ W; Mo) if spin-orbit coupling is neglected. We characterize the Dirac <span class="hlt">band</span> touching of these topological metals by theWilson loop of the non-Abelian Berry gauge field. Furthermore, we develop a criterion for the proximity of these topological metals to 2D and 3D Z 2 topological insulatorsmore » when spinorbit coupling is included; our criterion is based on nonsymmorphic symmetry eigenvalues, and may be used to identify topological materials without inversion symmetry. An <span class="hlt">additional</span> feature of the Dirac cone in monolayer MTe 2 is that it tilts over in a Lifshitz transition to produce electron and hole pockets—a type-II Dirac cone. These pockets, together with the pseudospin structure of the Dirac electrons, suggest a unified, topological explanation for the recently reported, nonsaturating magnetoresistance in WTe 2, as well as its circular dichroism in photoemission. We complement our analysis and first-principles <span class="hlt">band</span> structure calculations with an ab-initio-derived tight-binding model for the WTe 2 monolayer.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018PhyE...98..191L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PhyE...98..191L"><span>Discrete impurity <span class="hlt">band</span> from surface danging bonds in nitrogen and phosphorus doped SiC nanowires</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Li, Yan-Jing; Li, Shu-Long; Gong, Pei; Li, Ya-Lin; Cao, Mao-Sheng; Fang, Xiao-Yong</p> <p>2018-04-01</p> <p>The electronic structure and optical properties of the nitrogen and phosphorus doped silicon carbide nanowires (SiCNWs) are investigated using first-principle calculations based on density functional theory. The results show doping can change the type of the <span class="hlt">band</span> gap and improve the conductivity. However, the doped SiCNWs form a discrete impurity levels at the Fermi <span class="hlt">energy</span>, and the dispersion degree decreases with the diameter increasing. In order to reveal the root of this phenomenon, we hydrogenated the doped SiCNWs, found that the surface dangling bonds were saturated, and the discrete impurity levels are degeneracy, which indicates that the discrete impurity <span class="hlt">band</span> of the doped SiCNWs is derived from the dangling bonds. The surface passivation can degenerate the impurity levels. Therefore, both doping and surface passivation can better improve the photoelectric properties of the SiCNWs. The result can provide <span class="hlt">additional</span> candidates in producing nano-optoelectronic devices.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2010-title47-vol1/pdf/CFR-2010-title47-vol1-sec15-243.pdf','CFR'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2010-title47-vol1/pdf/CFR-2010-title47-vol1-sec15-243.pdf"><span>47 CFR 15.243 - Operation in the <span class="hlt">band</span> 890-940 MHz.</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collectionCfr.action?selectedYearFrom=2010&page.go=Go">Code of Federal Regulations, 2010 CFR</a></p> <p></p> <p>2010-10-01</p> <p>... 47 Telecommunication 1 2010-10-01 2010-10-01 false Operation in the <span class="hlt">band</span> 890-940 MHz. 15.243 Section 15.243 Telecommunication FEDERAL COMMUNICATIONS COMMISSION GENERAL RADIO FREQUENCY DEVICES Intentional Radiators Radiated Emission Limits, <span class="hlt">Additional</span> Provisions § 15.243 Operation in the <span class="hlt">band</span> 890-940...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2010-title47-vol1/pdf/CFR-2010-title47-vol1-sec15-241.pdf','CFR'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2010-title47-vol1/pdf/CFR-2010-title47-vol1-sec15-241.pdf"><span>47 CFR 15.241 - Operation in the <span class="hlt">band</span> 174-216 MHz.</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collectionCfr.action?selectedYearFrom=2010&page.go=Go">Code of Federal Regulations, 2010 CFR</a></p> <p></p> <p>2010-10-01</p> <p>... 47 Telecommunication 1 2010-10-01 2010-10-01 false Operation in the <span class="hlt">band</span> 174-216 MHz. 15.241 Section 15.241 Telecommunication FEDERAL COMMUNICATIONS COMMISSION GENERAL RADIO FREQUENCY DEVICES Intentional Radiators Radiated Emission Limits, <span class="hlt">Additional</span> Provisions § 15.241 Operation in the <span class="hlt">band</span> 174-216...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22584176-direct-optical-band-gap-measurement-polycrystalline-semiconductors-critical-look-tauc-method','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22584176-direct-optical-band-gap-measurement-polycrystalline-semiconductors-critical-look-tauc-method"><span>Direct optical <span class="hlt">band</span> gap measurement in polycrystalline semiconductors: A critical look at the Tauc method</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Dolgonos, Alex; Mason, Thomas O.; Poeppelmeier, Kenneth R., E-mail: krp@northwestern.edu</p> <p>2016-08-15</p> <p>The direct optical <span class="hlt">band</span> gap of semiconductors is traditionally measured by extrapolating the linear region of the square of the absorption curve to the x-axis, and a variation of this method, developed by Tauc, has also been widely used. The application of the Tauc method to crystalline materials is rooted in misconception–and traditional linear extrapolation methods are inappropriate for use on degenerate semiconductors, where the occupation of conduction <span class="hlt">band</span> <span class="hlt">energy</span> states cannot be ignored. A new method is proposed for extracting a direct optical <span class="hlt">band</span> gap from absorption spectra of degenerately-doped bulk semiconductors. This method was applied to pseudo-absorption spectramore » of Sn-doped In{sub 2}O{sub 3} (ITO)—converted from diffuse-reflectance measurements on bulk specimens. The results of this analysis were corroborated by room-temperature photoluminescence excitation measurements, which yielded values of optical <span class="hlt">band</span> gap and Burstein–Moss shift that are consistent with previous studies on In{sub 2}O{sub 3} single crystals and thin films. - Highlights: • The Tauc method of <span class="hlt">band</span> gap measurement is re-evaluated for crystalline materials. • Graphical method proposed for extracting optical <span class="hlt">band</span> gaps from absorption spectra. • The proposed method incorporates an <span class="hlt">energy</span> broadening term for <span class="hlt">energy</span> transitions. • Values for ITO were self-consistent between two different measurement methods.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JSP...170..399T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JSP...170..399T"><span>Ferromagnetism in the Hubbard Model with a Gapless Nearly-Flat <span class="hlt">Band</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Tanaka, Akinori</p> <p>2018-01-01</p> <p>We present a version of the Hubbard model with a gapless nearly-flat lowest <span class="hlt">band</span> which exhibits ferromagnetism in two or more dimensions. The model is defined on a lattice obtained by placing a site on each edge of the hypercubic lattice, and electron hopping is assumed to be only between nearest and next nearest neighbor sites. The lattice, where all the sites are identical, is simple, and the corresponding single-electron <span class="hlt">band</span> structure, where two cosine-type <span class="hlt">bands</span> touch without an <span class="hlt">energy</span> gap, is also simple. We prove that the ground state of the model is unique and ferromagnetic at half-filling of the lower <span class="hlt">band</span>, if the lower <span class="hlt">band</span> is nearly flat and the strength of on-site repulsion is larger than a certain value which is independent of the lattice size. This is the first example of ferromagnetism in three dimensional non-singular models with a gapless <span class="hlt">band</span> structure.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22907924','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22907924"><span>Clock is not a component of Z-<span class="hlt">bands</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wang, Jushuo; Dube, Dipak K; White, Jennifer; Fan, Yingli; Sanger, Jean M; Sanger, Joseph W</p> <p>2012-12-01</p> <p>The process of Z-<span class="hlt">band</span> assembly begins with the formation of small Z-bodies composed of a complex of proteins rich in alpha-actinin. As <span class="hlt">additional</span> proteins are added to nascent myofibrils, Z-bodies are transformed into continuous <span class="hlt">bands</span> that form coherent discs of interacting proteins at the boundaries of sarcomeres. The steps controlling the transition of Z-bodies to Z-<span class="hlt">bands</span> are not known. The report that a circadian protein, Clock, was localized in the Z-<span class="hlt">bands</span> of neonatal rat cardiomyocytes raised the question whether this transcription factor could be involved in Z-<span class="hlt">band</span> assembly. We found that the anti-Clock antibody used in the reported study also stained the Z-<span class="hlt">bands</span> and Z-bodies of mouse and avian cardiac and skeletal muscle cells. YFP constructs of Clock that were assembled, however, did not localize to the Z-<span class="hlt">bands</span> of muscle cells. Controls of Clock's activity showed that cotransfection of muscle cells with pYFP-Clock and pCeFP-BMAL1 led to the expected nuclear localization of YFP-Clock with its binding partner CeFP-BMAL1. Neither CeFP-BMAL1 nor antibodies directed against BMAL1 localized to Z-<span class="hlt">bands</span>. A bimolecular fluorescence complementation assay (VC-BMAL1 and VN-Clock) confirmed the absence of Clock and BMAL1 from Z-<span class="hlt">bands</span>, and their nuclear colocalization. A second anti-Clock antibody stained nuclei, but not Z-<span class="hlt">bands</span>, of cells cotransfected with Clock and BMAL1 plasmids. Western blots of reactions of muscle extracts and purified alpha-actinins with the two anti-Clock antibodies showed that the original antibody cross-reacted with alpha-actinin and the second did not. These results cannot confirm Clock as an active component of Z-<span class="hlt">bands</span>. © 2012 Wiley Periodicals, Inc. Copyright © 2012 Wiley Periodicals, Inc.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2003JAP....94.3931K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2003JAP....94.3931K"><span><span class="hlt">Band</span> line-up determination at p- and n-type Al/4H-SiC Schottky interfaces using photoemission spectroscopy</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kohlscheen, J.; Emirov, Y. N.; Beerbom, M. M.; Wolan, J. T.; Saddow, S. E.; Chung, G.; MacMillan, M. F.; Schlaf, R.</p> <p>2003-09-01</p> <p>The <span class="hlt">band</span> lineup of p- and n-type 4H-SiC/Al interfaces was determined using x-ray photoemission spectroscopy (XPS). Al was deposited in situ on ex situ cleaned SiC substrates in several steps starting at 1.2 Å up to 238 Å nominal film thickness. Before growth and after each growth step, the sample surface was characterized in situ by XPS. The analysis of the spectral shifts indicated that during the initial deposition stages the Al films react with the ambient surface contamination layer present on the samples after insertion into vacuum. At higher coverage metallic Al clusters are formed. The <span class="hlt">band</span> lineups were determined from the analysis of the core level peak shifts and the positions of the valence <span class="hlt">bands</span> maxima (VBM) depending on the Al overlayer thickness. Shifts of the Si 2p and C 1s XPS core levels occurred to higher (lower) binding <span class="hlt">energy</span> for the p-(n-)type substrates, which was attributed to the occurrence of <span class="hlt">band</span> bending due to Fermi-level equilibration at the interface. The hole injection barrier at the p-type interface was determined to be 1.83±0.1 eV, while the n-type interface revealed an electron injection barrier of 0.98±0.1 eV. Due to the weak features in the SiC valence <span class="hlt">bands</span> measured by XPS, the VBM positions were determined using the Si 2p peak positions. This procedure required the determination of the Si 2p-to-VBM binding <span class="hlt">energy</span> difference (99.34 eV), which was obtained from <span class="hlt">additional</span> measurements.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19970001766&hterms=cloud+database&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dcloud%2Bdatabase','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19970001766&hterms=cloud+database&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dcloud%2Bdatabase"><span>O2 A <span class="hlt">Band</span> Studies for Cloud Detection and Algorithm Improvement</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Chance, K. V.</p> <p>1996-01-01</p> <p>Detection of cloud parameters from space-based spectrometers can employ the vibrational <span class="hlt">bands</span> of O2 in the (sup b1)Sigma(sub +)(sub g) yields X(sub 3) Sigma(sup -)(sub g) spin-forbidden electronic transition manifold, particularly the Delta nu = 0 A <span class="hlt">band</span>. The GOME instrument uses the A <span class="hlt">band</span> in the Initial Cloud Fitting Algorithm (ICFA). The work reported here consists of making substantial improvements in the line-by-line spectral database for the A <span class="hlt">band</span>, testing whether an <span class="hlt">additional</span> correction to the line shape function is necessary in order to correctly model the atmospheric transmission in this <span class="hlt">band</span>, and calculating prototype cloud and ground template spectra for comparison with satellite measurements.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018SPIE10697E..13L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018SPIE10697E..13L"><span>The design and application of a multi-<span class="hlt">band</span> IR imager</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Li, Lijuan</p> <p>2018-02-01</p> <p>Multi-<span class="hlt">band</span> IR imaging system has many applications in security, national defense, petroleum and gas industry, etc. So the relevant technologies are getting more and more attention in rent years. As we know, when used in missile warning and missile seeker systems, multi-<span class="hlt">band</span> IR imaging technology has the advantage of high target recognition capability and low false alarm rate if suitable spectral <span class="hlt">bands</span> are selected. Compared with traditional single <span class="hlt">band</span> IR imager, multi-<span class="hlt">band</span> IR imager can make use of spectral features in <span class="hlt">addition</span> to space and time domain features to discriminate target from background clutters and decoys. So, one of the key work is to select the right spectral <span class="hlt">bands</span> in which the feature difference between target and false target is evident and is well utilized. Multi-<span class="hlt">band</span> IR imager is a useful instrument to collect multi-<span class="hlt">band</span> IR images of target, backgrounds and decoys for spectral <span class="hlt">band</span> selection study at low cost and with adjustable parameters and property compared with commercial imaging spectrometer. In this paper, a multi-<span class="hlt">band</span> IR imaging system is developed which is suitable to collect 4 spectral <span class="hlt">band</span> images of various scenes at every turn and can be expanded to other short-wave and mid-wave IR spectral <span class="hlt">bands</span> combination by changing filter groups. The multi-<span class="hlt">band</span> IR imaging system consists of a broad <span class="hlt">band</span> optical system, a cryogenic InSb large array detector, a spinning filter wheel and electronic processing system. The multi-<span class="hlt">band</span> IR imaging system's performance is tested in real data collection experiments.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017PhRvB..95o5310L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017PhRvB..95o5310L"><span>Understanding <span class="hlt">band</span> alignments in semiconductor heterostructures: Composition dependence and type-I-type-II transition of natural <span class="hlt">band</span> offsets in nonpolar zinc-blende AlxGa1 -xN /AlyGa1 -yN composites</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Landmann, M.; Rauls, E.; Schmidt, W. G.</p> <p>2017-04-01</p> <p>The composition dependence of the natural <span class="hlt">band</span> alignment at nonpolar AlxGa1 -xN /AlyGa1 -yN heterojunctions is investigated via hybrid functional based density functional theory. Accurate <span class="hlt">band</span>-gap data are provided using Heyd-Scuseria-Ernzerhof (HSE) type hybrid functionals with a composition dependent exact-exchange contribution. The unstrained <span class="hlt">band</span> alignment between zincblende (zb) AlxGa1 -xN semiconductor alloys is studied within the entire ternary composition range utilizing the Branch-point technique to align the <span class="hlt">energy</span> levels related to the bulklike direct Γv→Γc and indirect, pseudodirect, respectively, Γv→Xc type transitions in zb-AlxGa1 -xN . While the zb-GaN/AlxGa1 -xN <span class="hlt">band</span> edges consistently show a type-I alignment, the relative position of fundamental <span class="hlt">band</span> edges changes to a type-II alignment in the Al-rich composition ranges of zb-AlxGa1 -xN /AlN and zb-AlxGa1 -xN /AlyGa1 -yN systems. The presence of a direct-indirect <span class="hlt">band</span>-gap transition at xc=0.63 in zb-AlxGa1 -xN semiconductor alloys gives rise to a notably different composition dependence of <span class="hlt">band</span> discontinuities in the direct and indirect <span class="hlt">energy</span>-gap ranges. Below the critical direct-indirect Al/Ga-crossover concentration, the <span class="hlt">band</span> offsets show a close to linear dependence on the alloy composition. In contrast, notable bowing characteristics of all <span class="hlt">band</span> discontinuities are observed above the critical crossover composition.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22591250-flat-electronic-bands-fractal-kagome-network-effect-perturbation','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22591250-flat-electronic-bands-fractal-kagome-network-effect-perturbation"><span>Flat electronic <span class="hlt">bands</span> in fractal-kagomé network and the effect of perturbation</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Nandy, Atanu, E-mail: atanunandy1989@gmail.com; Chakrabarti, Arunava, E-mail: arunava-chakrabarti@yahoo.co.in</p> <p>2016-05-06</p> <p>We demonstrate an analytical prescription of demonstrating the flat <span class="hlt">band</span> [FB] states in a fractal incorporated kagomé type network that can give rise to a countable infinity of flat non-dispersive eigenstates with a multitude of localization area. The onset of localization can, in principle, be delayed in space by an appropriate choice of <span class="hlt">energy</span> regime. The length scale, at which the onset of localization for each mode occurs, can be tuned at will following the formalism developed within the framework of real space renormalization group. This scheme leads to an exact determination of <span class="hlt">energy</span> eigenvalue for which one can havemore » dispersionless flat electronic <span class="hlt">bands</span>. Furthermore, we have shown the effect ofuniform magnetic field for the same non-translationally invariant network model that has ultimately led to an‘apparent invisibility’ of such staggered localized states and to generate absolutely continuous sub-<span class="hlt">bands</span> in the <span class="hlt">energy</span> spectrum and again an interesting re-entrant behavior of those FB states.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28880064','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28880064"><span>Fullerene-Free Organic Solar Cells with an Efficiency of 10.2% and an <span class="hlt">Energy</span> Loss of 0.59 eV Based on a Thieno[3,4-c]Pyrrole-4,6-dione-Containing Wide <span class="hlt">Band</span> Gap Polymer Donor.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Hadmojo, Wisnu Tantyo; Wibowo, Febrian Tri Adhi; Ryu, Du Yeol; Jung, In Hwan; Jang, Sung-Yeon</p> <p>2017-09-27</p> <p>Although the combination of wide <span class="hlt">band</span> gap polymer donors and narrow <span class="hlt">band</span> gap small-molecule acceptors achieved state-of-the-art performance as bulk heterojunction (BHJ) active layers for organic solar cells, there have been only several of the wide <span class="hlt">band</span> gap polymers that actually realized high-efficiency devices over >10%. Herein, we developed high-efficiency, low-<span class="hlt">energy</span>-loss fullerene-free organic solar cells using a weakly crystalline wide <span class="hlt">band</span> gap polymer donor, PBDTTPD-HT, and a nonfullerene small-molecule acceptor, ITIC. The excessive intermolecular stacking of ITIC is efficiently suppressed by the miscibility with PBDTTPD-HT, which led to a well-balanced nanomorphology in the PBDTTPD-HT/ITIC BHJ active films. The favorable optical, electronic, and energetic properties of PBDTTPD-HT with respect to ITIC achieved panchromatic photon-to-current conversion with a remarkably low <span class="hlt">energy</span> loss (0.59 eV).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/15195466','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/15195466"><span>Assessment of the influence of <span class="hlt">energy</span> under-reporting on intake estimates of four food <span class="hlt">additives</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Gilsenan, M B; Gibney, M J</p> <p>2004-03-01</p> <p>Under-reporting has been identified as an important source of uncertainty in food chemical exposure assessments. The objective of the present study was to assess the influence of under-reporting on food <span class="hlt">additive</span> intake estimates. Dietary survey data were derived from the North-South Ireland Food Consumption Survey (2001). Data from the Republic of Ireland (n = 958) were used. <span class="hlt">Energy</span> under-reporters were identified using a ratio of <span class="hlt">energy</span> intakes to estimated basal metabolic rate. First, food categories (n = 26) included in an assessment of exposure of four food <span class="hlt">additives</span> were created and patterns of food intakes (i.e. likelihood of consumption, frequency of consumption and reported portion size) between acceptable and under-reporters compared. Second, for each food <span class="hlt">additive</span>, deterministic intake estimates for the total sample (i.e. acceptable and under-reporters), under-reporters and acceptable reporters were calculated and compared. Differential reporting of the majority of food categories between acceptable and under-reporters was recorded. Under-reporters were less likely to record the consumption of a given food and more likely to under-report the frequency of consumption and portion size compared with acceptable reporters. Food <span class="hlt">additive</span> intake estimates amongst acceptable reporters were higher than corresponding intake estimates amongst the total sample of reporters and amongst under-reporters. With the exception of one food <span class="hlt">additive</span> (erythrosine), ratios of upper percentile <span class="hlt">additive</span> intakes amongst acceptable reporters to corresponding intake estimates amongst the total sample of reporters did not exceed 1.06 when results were expressed as total population or consumer-only intakes. Findings illustrated that <span class="hlt">energy</span> under-reporting does not materially influence estimates of food <span class="hlt">additive</span> exposure based on the four food <span class="hlt">additives</span> studied. However, a number of situations were identified where the under-reporting might exert a more significant impact on resulting</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015APS..MAR.G9001M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015APS..MAR.G9001M"><span>Theory of <span class="hlt">Band</span> Warping and its Effects on Thermoelectronic Transport Properties</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mecholsky, Nicholas; Resca, Lorenzo; Pegg, Ian; Fornari, Marco</p> <p>2015-03-01</p> <p>Transport properties of materials depend upon features of <span class="hlt">band</span> structures near extrema in the BZ. Such features are generally described in terms of quadratic expansions and effective masses. Such expansions, however, are permissible only under strict conditions that are sometimes violated by materials. Suggestive terms such as ``<span class="hlt">band</span> warping'' have been used to refer to such situations and ad hoc methods have been developed to treat them. We develop a generally applicable theory, based on radial expansions, and a corresponding definition of angular effective mass which also accounts for effects of <span class="hlt">band</span> non-parabolicity and anisotropy. Further, we develop precise procedures to evaluate <span class="hlt">band</span> warping quantitatively and as an example we analyze the warping features of valence <span class="hlt">bands</span> in silicon using first-principles calculations and we compare those with semi-empirical models. We use our theory to generalize derivations of transport coefficients for cases of either single or multiple electronic <span class="hlt">bands</span>, with either quadratically expansible or warped <span class="hlt">energy</span> surfaces. We introduce the transport-equivalent ellipsoid and illustrate the drastic effects that <span class="hlt">band</span> warping can induce on thermoelectric properties using multi-<span class="hlt">band</span> models. Vitreous State Laboratory and Samsung's GRO program.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1993SPIE.1941..221H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1993SPIE.1941..221H"><span>Comparison of C-<span class="hlt">band</span> and Ku-<span class="hlt">band</span> scatterometry for medium-resolution tropical forest inventory</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hardin, Perry J.; Long, David G.</p> <p>1993-08-01</p> <p>Since 1978, AVHRR imagery from NOAA polar orbiters has provided coverage of tropical regions at this desirable resolution, but much of the imagery is plagued with heavy cloud cover typical of equatorial regions. Clearly a medium resolution radar sensor would be a useful <span class="hlt">addition</span> to AVHRR, but none are planned to fly in the future. In contrast, scatterometers are an important radar component of many future earth remote sensing systems, but the inherent resolution of these instruments is too low (approximately equals 50 km) for monitoring earth's land surfaces. However, a recently developed image reconstruction technique can increase the spatial resolution of scatterometer data to levels (approximately equals 4 to 14 km) approaching AVHRR global area coverage (approximately equals 4 km). When reconstructed, scatterometer data may prove to be an important asset in evaluating equatorial land cover. In this paper, the authors compare the utility of reconstructed Seasat scatterometer (SASS), Ku-<span class="hlt">band</span> microwave data to reconstructed ERS-1 C-<span class="hlt">band</span> scatterometer imagery for discrimination and monitoring of tropical vegetation formations. In comparative classification experiments conducted on reconstructed images of Brasil, the ERS-1 C-<span class="hlt">band</span> imagery was slightly superior to its reconstructed SASS Ku-<span class="hlt">band</span> counterpart for discriminating between several equatorial land cover classes. A classification accuracy approaching .90 was achieved when the two scatterometer images were combined with an AVHRR normalized difference vegetation index (NDVI) image. The success of these experiments indicates that further research into reconstructed image applications to tropical forest monitoring is warranted.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5025745','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5025745"><span>Madelung and Hubbard interactions in polaron <span class="hlt">band</span> model of doped organic semiconductors</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Png, Rui-Qi; Ang, Mervin C.Y.; Teo, Meng-How; Choo, Kim-Kian; Tang, Cindy Guanyu; Belaineh, Dagmawi; Chua, Lay-Lay; Ho, Peter K.H.</p> <p>2016-01-01</p> <p>The standard polaron <span class="hlt">band</span> model of doped organic semiconductors predicts that density-of-states shift into the π–π* gap to give a partially filled polaron <span class="hlt">band</span> that pins the Fermi level. This picture neglects both Madelung and Hubbard interactions. Here we show using ultrahigh workfunction hole-doped model triarylamine–fluorene copolymers that Hubbard interaction strongly splits the singly-occupied molecular orbital from its empty counterpart, while Madelung (Coulomb) interactions with counter-anions and other carriers markedly shift <span class="hlt">energies</span> of the frontier orbitals. These interactions lower the singly-occupied molecular orbital <span class="hlt">band</span> below the valence <span class="hlt">band</span> edge and give rise to an empty low-lying counterpart <span class="hlt">band</span>. The Fermi level, and hence workfunction, is determined by conjunction of the bottom edge of this empty <span class="hlt">band</span> and the top edge of the valence <span class="hlt">band</span>. Calculations are consistent with the observed Fermi-level downshift with counter-anion size and the observed dependence of workfunction on doping level in the strongly doped regime. PMID:27582355</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li class="active"><span>21</span></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_21 --> <div id="page_22" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li class="active"><span>22</span></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li><a href="#" onclick='return showDiv("page_24");'>24</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="421"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27421066','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27421066"><span>Narrow <span class="hlt">Band</span> Gap Lead Sulfide Hole Transport Layers for Quantum Dot Photovoltaics.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zhang, Nanlin; Neo, Darren C J; Tazawa, Yujiro; Li, Xiuting; Assender, Hazel E; Compton, Richard G; Watt, Andrew A R</p> <p>2016-08-24</p> <p>The <span class="hlt">band</span> structure of colloidal quantum dot (CQD) bilayer heterojunction solar cells is optimized using a combination of ligand modification and QD <span class="hlt">band</span> gap control. Solar cells with power conversion efficiencies of up to 9.33 ± 0.50% are demonstrated by aligning the absorber and hole transport layers (HTL). Key to achieving high efficiencies is optimizing the relative position of both the valence <span class="hlt">band</span> and Fermi <span class="hlt">energy</span> at the CQD bilayer interface. By comparing different <span class="hlt">band</span> gap CQDs with different ligands, we find that a smaller <span class="hlt">band</span> gap CQD HTL in combination with a more p-type-inducing CQD ligand is found to enhance hole extraction and hence device performance. We postulate that the efficiency improvements observed are largely due to the synergistic effects of narrower <span class="hlt">band</span> gap QDs, causing an upshift of valence <span class="hlt">band</span> position due to 1,2-ethanedithiol (EDT) ligands and a lowering of the Fermi level due to oxidation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.jstor.org/stable/3798230','USGSPUBS'); return false;" href="http://www.jstor.org/stable/3798230"><span>Reward <span class="hlt">banding</span> to determine reporting rate of recovered mourning dove <span class="hlt">bands</span></span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Tomlinson, R.E.</p> <p>1968-01-01</p> <p>Reward <span class="hlt">bands</span> placed on the other leg of certain regularly <span class="hlt">banded</span> immature mourning doves (Zenaidura macroura) were used to develop information on reporting rates of recovered dove <span class="hlt">bands</span>. Reports from 15 widely separated sections of the United States showed considerable variation in recovery rate of doves both with and without reward <span class="hlt">bands</span>. The overall percentages of <span class="hlt">banded</span> doves that were reported as recovered were 9.69% for those with reward <span class="hlt">bands</span> and 3.83% for controls. The bandreporting rate for states influenced by publicity was 66%; that for states not influenced was 32%.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AnGeo..35.1069P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AnGeo..35.1069P"><span>Statistical study of auroral omega <span class="hlt">bands</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Partamies, Noora; Weygand, James M.; Juusola, Liisa</p> <p>2017-09-01</p> <p>The presence of very few statistical studies on auroral omega <span class="hlt">bands</span> motivated us to test-use a semi-automatic method for identifying large-scale undulations of the diffuse aurora boundary and to investigate their occurrence. Five identical all-sky cameras with overlapping fields of view provided data for 438 auroral omega-like structures over Fennoscandian Lapland from 1996 to 2007. The results from this set of omega <span class="hlt">band</span> events agree remarkably well with previous observations of omega <span class="hlt">band</span> occurrence in magnetic local time (MLT), lifetime, location between the region 1 and 2 field-aligned currents, as well as current density estimates. The average peak emission height of omega forms corresponds to the estimated precipitation <span class="hlt">energies</span> of a few keV, which experienced no significant change during the events. Analysis of both local and global magnetic indices demonstrates that omega <span class="hlt">bands</span> are observed during substorm expansion and recovery phases that are more intense than average substorm expansion and recovery phases in the same region. The omega occurrence with respect to the substorm expansion and recovery phases is in a very good agreement with an earlier observed distribution of fast earthward flows in the plasma sheet during expansion and recovery phases. These findings support the theory that omegas are produced by fast earthward flows and auroral streamers, despite the rarity of good conjugate observations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014PhRvB..90g5203F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014PhRvB..90g5203F"><span><span class="hlt">Band</span> gap renormalization and Burstein-Moss effect in silicon- and germanium-doped wurtzite GaN up to 1020 cm-3</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Feneberg, Martin; Osterburg, Sarah; Lange, Karsten; Lidig, Christian; Garke, Bernd; Goldhahn, Rüdiger; Richter, Eberhard; Netzel, Carsten; Neumann, Maciej D.; Esser, Norbert; Fritze, Stephanie; Witte, Hartmut; Bläsing, Jürgen; Dadgar, Armin; Krost, Alois</p> <p>2014-08-01</p> <p>The interplay between <span class="hlt">band</span> gap renormalization and <span class="hlt">band</span> filling (Burstein-Moss effect) in n-type wurtzite GaN is investigated. For a wide range of electron concentrations up to 1.6×1020cm-3 spectroscopic ellipsometry and photoluminescence were used to determine the dependence of the <span class="hlt">band</span> gap <span class="hlt">energy</span> and the Fermi edge on electron density. The <span class="hlt">band</span> gap renormalization is the dominating effect up to an electron density of about 9×1018cm-3; at higher values the Burstein-Moss effect is stronger. Exciton screening, the Mott transition, and formation of Mahan excitons are discussed. A quantitative understanding of the near gap transition <span class="hlt">energies</span> on electron density is obtained. Higher <span class="hlt">energy</span> features in the dielectric functions up to 10eV are not influenced by <span class="hlt">band</span> gap renormalization.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011RScI...82c4704L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011RScI...82c4704L"><span>Electrically detected magnetic resonance in a W-<span class="hlt">band</span> microwave cavity</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lang, V.; Lo, C. C.; George, R. E.; Lyon, S. A.; Bokor, J.; Schenkel, T.; Ardavan, A.; Morton, J. J. L.</p> <p>2011-03-01</p> <p>We describe a low-temperature sample probe for the electrical detection of magnetic resonance in a resonant W-<span class="hlt">band</span> (94 GHz) microwave cavity. The advantages of this approach are demonstrated by experiments on silicon field-effect transistors. A comparison with conventional low-frequency measurements at X-<span class="hlt">band</span> (9.7 GHz) on the same devices reveals an up to 100-fold enhancement of the signal intensity. In <span class="hlt">addition</span>, resonance lines that are unresolved at X-<span class="hlt">band</span> are clearly separated in the W-<span class="hlt">band</span> measurements. Electrically detected magnetic resonance at high magnetic fields and high microwave frequencies is therefore a very sensitive technique for studying electron spins with an enhanced spectral resolution and sensitivity.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20170005641','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20170005641"><span>Satellite Communications for Unmanned Aircraft C2 Links: C-<span class="hlt">Band</span>, Ku-<span class="hlt">Band</span> and Ka-<span class="hlt">Band</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Kerczewski, Robert J.; Wilson, Jeffrey D.; Bishop, William D.</p> <p>2016-01-01</p> <p>Unmanned aircraft (UA) that require access to controlled (or non-segregated) airspace require a highly reliable and robust command and control (C2) link, operating over protected aviation spectrum. While operating within radio line-of-sight (LOS) UA can make use of air-to-ground C2 links to terrestrial stations. When operating beyond LOS (BLOS) where a group of networked terrestrial stations does not exist to provide effective BLOS coverage, a satellite communications link is required. Protected aviation spectrum for satellite C2 links has only recently been allocated in <span class="hlt">bands</span> where operational satellites exist. A previously existing C-<span class="hlt">Band</span> allocation covers a <span class="hlt">bands</span> where there are currently no operational satellites. The new allocations, within the Fixed Satellite Service <span class="hlt">bands</span> at Ku and Ka-<span class="hlt">Bands</span> will not be finalized until 2023 due to the need for the development of standards and technical decisions on the operation of UA satellite C2 links within these <span class="hlt">bands</span>. This paper provides an overview of BLOS satellite C2 links, some of the conditions which will need to be met for the operation of such links, and a look at some aspects of spectrum sharing which may constrain these operations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22667271-high-energy-electron-irradiation-interstellar-carbonaceous-dust-analogs-cosmic-ray-effects-carriers-absorption-band','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22667271-high-energy-electron-irradiation-interstellar-carbonaceous-dust-analogs-cosmic-ray-effects-carriers-absorption-band"><span>HIGH-<span class="hlt">ENERGY</span> ELECTRON IRRADIATION OF INTERSTELLAR CARBONACEOUS DUST ANALOGS: COSMIC-RAY EFFECTS ON THE CARRIERS OF THE 3.4 μ m ABSORPTION <span class="hlt">BAND</span></span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Maté, Belén; Molpeceres, Germán; Jiménez-Redondo, Miguel</p> <p>2016-11-01</p> <p>The effects of cosmic rays on the carriers of the interstellar 3.4 μ m absorption <span class="hlt">band</span> have been investigated in the laboratory. This <span class="hlt">band</span> is attributed to stretching vibrations of CH{sub 3} and CH{sub 2} in carbonaceous dust. It is widely observed in the diffuse interstellar medium, but disappears in dense clouds. Destruction of CH{sub 3} and CH{sub 2} by cosmic rays could become relevant in dense clouds, shielded from the external ultraviolet field. For the simulations, samples of hydrogenated amorphous carbon (a-C:H) have been irradiated with 5 keV electrons. The decay of the <span class="hlt">band</span> intensity versus electron fluence reflectsmore » a-C:H dehydrogenation, which is well described by a model assuming that H{sub 2} molecules, formed by the recombination of H atoms liberated through CH bond breaking, diffuse out of the sample. The CH bond destruction rates derived from the present experiments are in good accordance with those from previous ion irradiation experiments of HAC. The experimental simplicity of electron bombardment has allowed the use of higher-<span class="hlt">energy</span> doses than in the ion experiments. The effects of cosmic rays on the aliphatic components of cosmic dust are found to be small. The estimated cosmic-ray destruction times for the 3.4 μ m <span class="hlt">band</span> carriers lie in the 10{sup 8} yr range and cannot account for the disappearance of this <span class="hlt">band</span> in dense clouds, which have characteristic lifetimes of 3 × 10{sup 7} yr. The results invite a more detailed investigation of the mechanisms of CH bond formation and breaking in the intermediate region between diffuse and dense clouds.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24734845','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24734845"><span>Secular trends in Cherokee cranial morphology: Eastern vs Western <span class="hlt">bands</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Sutphin, Rebecca; Ross, Ann H; Jantz, Richard L</p> <p>2014-01-01</p> <p>The research objective was to examine if secular trends can be identified for cranial data commissioned by Boas in 1892, specifically for cranial breadth and cranial length of the Eastern and Western <span class="hlt">band</span> Cherokee who experienced environmental hardships. Multiple regression analysis was used to test the degree of relationship between each of the cranial measures: cranial length, cranial breadth and cephalic index, along with predictor variables (year-of-birth, location, sex, admixture); the model revealed a significant difference for all craniometric variables. <span class="hlt">Additional</span> regression analysis was performed with smoothing Loess plots to observe cranial length and cranial breadth change over time (year-of-birth) separately for Eastern and Western Cherokee <span class="hlt">band</span> females and males born between 1783-1874. This revealed the Western and Eastern <span class="hlt">bands</span> show a decrease in cranial length over time. Eastern <span class="hlt">band</span> individuals maintain a relatively constant head breadth, while Western <span class="hlt">Band</span> individuals show a sharp decline beginning around 1860. These findings support negative secular trend occurring for both Cherokee <span class="hlt">bands</span> where the environment made a detrimental impact; this is especially marked with the Western Cherokee <span class="hlt">band</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29757801','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29757801"><span>Effects of <span class="hlt">Additional</span> Low-Pass-Filtered Speech on Listening Effort for Noise-<span class="hlt">Band</span>-Vocoded Speech in Quiet and in Noise.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Pals, Carina; Sarampalis, Anastasios; van Dijk, Mart; Başkent, Deniz</p> <p>2018-05-11</p> <p>Residual acoustic hearing in electric-acoustic stimulation (EAS) can benefit cochlear implant (CI) users in increased sound quality, speech intelligibility, and improved tolerance to noise. The goal of this study was to investigate whether the low-pass-filtered acoustic speech in simulated EAS can provide the <span class="hlt">additional</span> benefit of reducing listening effort for the spectrotemporally degraded signal of noise-<span class="hlt">band</span>-vocoded speech. Listening effort was investigated using a dual-task paradigm as a behavioral measure, and the NASA Task Load indeX as a subjective self-report measure. The primary task of the dual-task paradigm was identification of sentences presented in three experiments at three fixed intelligibility levels: at near-ceiling, 50%, and 79% intelligibility, achieved by manipulating the presence and level of speech-shaped noise in the background. Listening effort for the primary intelligibility task was reflected in the performance on the secondary, visual response time task. Experimental speech processing conditions included monaural or binaural vocoder, with added low-pass-filtered speech (to simulate EAS) or without (to simulate CI). In Experiment 1, in quiet with intelligibility near-ceiling, <span class="hlt">additional</span> low-pass-filtered speech reduced listening effort compared with binaural vocoder, in line with our expectations, although not compared with monaural vocoder. In Experiments 2 and 3, for speech in noise, added low-pass-filtered speech allowed the desired intelligibility levels to be reached at less favorable speech-to-noise ratios, as expected. It is interesting that this came without the cost of increased listening effort usually associated with poor speech-to-noise ratios; at 50% intelligibility, even a reduction in listening effort on top of the increased tolerance to noise was observed. The NASA Task Load indeX did not capture these differences. The dual-task results provide partial evidence for a potential decrease in listening effort as a result of</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28269286','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28269286"><span>Alpha-<span class="hlt">band</span> rhythm suppression during memory recall reflecting memory performance.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Yokosawa, Koichi; Kimura, Keisuke; Chitose, Ryota; Momiki, Takuya; Kuriki, Shinya</p> <p>2016-08-01</p> <p>Alpha-<span class="hlt">band</span> rhythm is thought to be involved in memory processes, similarly to other spontaneous brain rhythms. Ten right-handed healthy volunteers participated in our proposed sequential short-term memory task that provides a serial position effect in accuracy rate. We recorded alpha-<span class="hlt">band</span> rhythms by magnetoencephalography during performance of the task and observed that the amplitude of the rhythm was suppressed dramatically in the memory recall period. The suppressed region was estimated to be in the occipital lobe, suggesting that alpha-<span class="hlt">band</span> rhythm is suppressed by activation of the occipital attentional network. <span class="hlt">Additionally</span>, the alpha-<span class="hlt">band</span> suppression reflected accuracy rate, that is, the amplitude was suppressed more when recalling items with higher accuracy rate. The sensors with a significant correlation between alpha-<span class="hlt">band</span> amplitude and accuracy rate were located widely from the frontal to occipital regions mainly in the right hemisphere. The results suggests that alpha-<span class="hlt">band</span> rhythm is involved in memory recall and can be index of memory performance.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018SPIE10607E..0TY','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018SPIE10607E..0TY"><span>An enhanced narrow-<span class="hlt">band</span> imaging method for the microvessel detection</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yu, Feng; Song, Enmin; Liu, Hong; Wan, Youming; Zhu, Jun; Hung, Chih-Cheng</p> <p>2018-02-01</p> <p>A medical endoscope system combined with the narrow-<span class="hlt">band</span> imaging (NBI), has been shown to be a superior diagnostic tool for early cancer detection. The NBI can reveal the morphologic changes of microvessels in the superficial cancer. In order to improve the conspicuousness of microvessel texture, we propose an enhanced NBI method to improve the conspicuousness of endoscopic images. To obtain the more conspicuous narrow-<span class="hlt">band</span> images, we use the edge operator to extract the edge information of the narrow-<span class="hlt">band</span> blue and green images, and give a weight to the extracted edges. Then, the weighted edges are fused with the narrow-<span class="hlt">band</span> blue and green images. Finally, the displayed endoscopic images are reconstructed with the enhanced narrow-<span class="hlt">band</span> images. In <span class="hlt">addition</span>, we evaluate the performance of enhanced narrow-<span class="hlt">band</span> images with different edge operators. Experimental results indicate that the Sobel and Canny operators achieve the best performance of all. Compared with traditional NBI method of Olympus company, our proposed method has more conspicuous texture of microvessel.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1427630','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1427630"><span><span class="hlt">Additively</span> Manufactured, Net Shape Powder Metallurgy Cans for Valves Used in <span class="hlt">Energy</span> Production</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Peter, William H.; Gandy, David; Lannom, Robert</p> <p></p> <p>This CRADA NFE-14-05241 was conducted as a Technical Collaboration project within the Oak Ridge National Laboratory (ORNL) Manufacturing Demonstration Facility (MDF) sponsored by the US Department of <span class="hlt">Energy</span> Advanced Manufacturing Office (CPS Agreement Number 24761). Opportunities for MDF technical collaborations are listed in the announcement “Manufacturing Demonstration Facility Technology Collaborations for US Manufacturers in Advanced Manufacturing and Materials Technologies” posted at http://web.ornl.gov/sci/manufacturing/docs/FBO-ORNL-MDF-2013-2.pdf. The goal of technical collaborations is to engage industry partners to participate in short-term, collaborative projects within the Manufacturing Demonstration Facility (MDF) to assess applicability and of new <span class="hlt">energy</span> efficient manufacturing technologies. Research sponsored by the U.S. Departmentmore » of <span class="hlt">Energy</span>, Office of <span class="hlt">Energy</span> Efficiency and Renewable <span class="hlt">Energy</span>, Advanced Manufacturing Office, under contract DE-AC05-00OR22725 with UT-Battelle, LLC.ORNL would like to acknowledge the leadership of EPRI in pulling together the extensive team and managing the execution of the project. In <span class="hlt">addition</span>, ORNL would like to acknowledge the other contributions of the team members associated with this project. Quintus provided time, access, expertise, and labor of their hydro forming capabilities to evaluate both conventional and <span class="hlt">additively</span> manufactured tools through this process. Crane ChemPharma <span class="hlt">Energy</span> provided guidance and information on valve geometries. Carpenter Powder Products was involved with the team providing information on powder processing as it pertains to the canning and hot isostatic pressing of powder. on providing powder and knowledge as it pertains to powder supply for hot isostatic pressing; they also provided powder for the test trials by the industrial team. Bodycote provided guidance on hot isostatic pressing and can requirements. They were also responsible for the hot isostatic pressing of the test</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012JPhCS.377a2093S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012JPhCS.377a2093S"><span>Electron <span class="hlt">band</span> structure of the high pressure cubic phase of AlH3</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Shi, Hongliang; Zarifi, Niliffar; Yim, Wai-Leung; Tse, J. S.</p> <p>2012-07-01</p> <p>The electronic <span class="hlt">band</span> structure of the cubic Pm3n phase of AlH3 stable above 100 GPa is examined with semi-local, Tran-Blaha modified Becke-Johnson local density approximation (TB-mBJLDA), screened hybrid density functionals and GW methods. The shift of the conduction <span class="hlt">band</span> to higher <span class="hlt">energy</span> with increasing pressure is predicted by all methods. However, there are significant differences in detail <span class="hlt">band</span> structure. In the pressure range from 90 to160 GPa, semi-local, hybrid functional and TB-mBJLDA calculations predicted that AlH3 is a poor metal. In comparison, GW calculations show a gap opening at 160 GPa and AlH3 becomes a small gap semi-conductor. From the trends of the calculated <span class="hlt">band</span> shifts, it can be concluded that the favourable conditions leading to the nesting of Fermi surfaces predicted by semi-local calculation have disappeared if the exchange term is included. The results highlight the importance of the correction to the exchange <span class="hlt">energy</span> on the <span class="hlt">band</span> structure of hydrogen dominant dense metal hydrides at high pressure hydrides and may help to rationalize the absence of superconductivity in AlH3 from experimental measurements.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1991ApPhL..58.2126S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1991ApPhL..58.2126S"><span>Valence-<span class="hlt">band</span>-edge shift due to doping in p + GaAs</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Silberman, J. A.; de Lyon, T. J.; Woodall, J. M.</p> <p>1991-05-01</p> <p>Accurate knowledge of the shifts in valence- and conduction-<span class="hlt">band</span> edges due to heavy doping effects is crucial in modeling GaAs device structures that utilize heavily doped layers. X-ray photoemission spectroscopy was used to deduce the shift in the valence-<span class="hlt">band</span>-edge induced by carbon (p type) doping to a carrier density of 1×1020 cm-3 based on a determination of the bulk binding <span class="hlt">energy</span> of the Ga and As core levels in this material. Analysis of the data indicates that the shift of the valence-<span class="hlt">band</span> maximum into the gap and the penetration of the Fermi level into the valence <span class="hlt">bands</span> exactly compensate at this degenerate carrier concentration, to give ΔEv =0.12±0.05 eV.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20120010450','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20120010450"><span>Dichroic Filter for Separating W-<span class="hlt">Band</span> and Ka-<span class="hlt">Band</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Epp, Larry W.; Durden, Stephen L.; Jamnejad, Vahraz; Long, Ezra M.; Sosnowski, John B.; Higuera, Raymond J.; Chen, Jacqueline C.</p> <p>2012-01-01</p> <p>The proposed Aerosol/Cloud/Ecosystems (ACEs) mission development would advance cloud profiling radar from that used in CloudSat by adding a 35-GHz (Ka-<span class="hlt">band</span>) channel to the 94-GHz (W-<span class="hlt">band</span>) channel used in CloudSat. In order to illuminate a single antenna, and use CloudSat-like quasi-optical transmission lines, a spatial diplexer is needed to add the Ka-<span class="hlt">band</span> channel. A dichroic filter separates Ka-<span class="hlt">band</span> from W-<span class="hlt">band</span> by employing advances in electrical discharge machining (EDM) and mode-matching analysis techniques developed and validated for designing dichroics for the Deep Space Network (DSN), to develop a preliminary design that both met the requirements of frequency separation and mechanical strength. First, a mechanical prototype was built using an approximately 102-micron-diameter EDM process, and tolerances of the hole dimensions, wall thickness, radius, and dichroic filter thickness measured. The prototype validated the manufacturing needed to design a dichroic filter for a higher-frequency usage than previously used in the DSN. The initial design was based on a Ka-<span class="hlt">band</span> design, but thicker walls are required for mechanical rigidity than one obtains by simply scaling the Ka-<span class="hlt">band</span> dichroic filter. The resulting trade of hole dimensions for mechanical rigidity (wall thickness) required electrical redesign of the hole dimensions. Updates to existing codes in the linear solver decreased the analysis time using mode-matching, enabling the electrical design to be realized quickly. This work is applicable to missions and instruments that seek to extend W-<span class="hlt">band</span> cloud profiling measurements to other frequencies. By demonstrating a dichroic filter that passes W-<span class="hlt">band</span>, but reflects a lower frequency, this opens up the development of instruments that both compare to and enhance CloudSat.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24283411','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24283411"><span>The dependence of graphene Raman D-<span class="hlt">band</span> on carrier density.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Liu, Junku; Li, Qunqing; Zou, Yuan; Qian, Qingkai; Jin, Yuanhao; Li, Guanhong; Jiang, Kaili; Fan, Shoushan</p> <p>2013-01-01</p> <p>Raman spectroscopy has been an integral part of graphene research and can provide information about graphene structure, electronic characteristics, and electron-phonon interactions. In this study, the characteristics of the graphene Raman D-<span class="hlt">band</span>, which vary with carrier density, are studied in detail, including the frequency, full width half-maximum, and intensity. We find the Raman D-<span class="hlt">band</span> frequency increases for hole doping and decreases for electron doping. The Raman D-<span class="hlt">band</span> intensity increases when the Fermi level approaches half of the excitation <span class="hlt">energy</span> and is higher in the case of electron doping than that of hole doping. These variations can be explained by electron-phonon interaction theory and quantum interference between different Raman pathways in graphene. The intensity ratio of Raman D- and G-<span class="hlt">band</span>, which is important for defects characterization in graphene, shows a strong dependence on carrier density.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5855439','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5855439"><span>The Effects of Graphene Stacking on the Performance of Methane Sensor: A First-Principles Study on the Adsorption, <span class="hlt">Band</span> Gap and Doping of Graphene</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Yang, Daoguo; Zhang, Guoqi; Chen, Liangbiao; Liu, Dongjing; Cai, Miao; Fan, Xuejun</p> <p>2018-01-01</p> <p>The effects of graphene stacking are investigated by comparing the results of methane adsorption <span class="hlt">energy</span>, electronic performance, and the doping feasibility of five dopants (i.e., B, N, Al, Si, and P) via first-principles theory. Both zigzag and armchair graphenes are considered. It is found that the zigzag graphene with Bernal stacking has the largest adsorption <span class="hlt">energy</span> on methane, while the armchair graphene with Order stacking is opposite. In <span class="hlt">addition</span>, both the Order and Bernal stacked graphenes possess a positive linear relationship between adsorption <span class="hlt">energy</span> and layer number. Furthermore, they always have larger adsorption <span class="hlt">energy</span> in zigzag graphene. For electronic properties, the results show that the stacking effects on <span class="hlt">band</span> gap are significant, but it does not cause big changes to <span class="hlt">band</span> structure and density of states. In the comparison of distance, the average interlamellar spacing of the Order stacked graphene is the largest. Moreover, the adsorption effect is the result of the interactions between graphene and methane combined with the change of graphene’s structure. Lastly, the armchair graphene with Order stacking possesses the lowest formation <span class="hlt">energy</span> in these five dopants. It could be the best choice for doping to improve the methane adsorption. PMID:29389860</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29389860','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29389860"><span>The Effects of Graphene Stacking on the Performance of Methane Sensor: A First-Principles Study on the Adsorption, <span class="hlt">Band</span> Gap and Doping of Graphene.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Yang, Ning; Yang, Daoguo; Zhang, Guoqi; Chen, Liangbiao; Liu, Dongjing; Cai, Miao; Fan, Xuejun</p> <p>2018-02-01</p> <p>The effects of graphene stacking are investigated by comparing the results of methane adsorption <span class="hlt">energy</span>, electronic performance, and the doping feasibility of five dopants (i.e., B, N, Al, Si, and P) via first-principles theory. Both zigzag and armchair graphenes are considered. It is found that the zigzag graphene with Bernal stacking has the largest adsorption <span class="hlt">energy</span> on methane, while the armchair graphene with Order stacking is opposite. In <span class="hlt">addition</span>, both the Order and Bernal stacked graphenes possess a positive linear relationship between adsorption <span class="hlt">energy</span> and layer number. Furthermore, they always have larger adsorption <span class="hlt">energy</span> in zigzag graphene. For electronic properties, the results show that the stacking effects on <span class="hlt">band</span> gap are significant, but it does not cause big changes to <span class="hlt">band</span> structure and density of states. In the comparison of distance, the average interlamellar spacing of the Order stacked graphene is the largest. Moreover, the adsorption effect is the result of the interactions between graphene and methane combined with the change of graphene's structure. Lastly, the armchair graphene with Order stacking possesses the lowest formation <span class="hlt">energy</span> in these five dopants. It could be the best choice for doping to improve the methane adsorption.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2010-title47-vol1/pdf/CFR-2010-title47-vol1-sec15-239.pdf','CFR'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2010-title47-vol1/pdf/CFR-2010-title47-vol1-sec15-239.pdf"><span>47 CFR 15.239 - Operation in the <span class="hlt">band</span> 88-108 MHz.</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collectionCfr.action?selectedYearFrom=2010&page.go=Go">Code of Federal Regulations, 2010 CFR</a></p> <p></p> <p>2010-10-01</p> <p>... Intentional Radiators Radiated Emission Limits, <span class="hlt">Additional</span> Provisions § 15.239 Operation in the <span class="hlt">band</span> 88-108 MHz. (a) Emissions from the intentional radiator shall be confined within a <span class="hlt">band</span> 200 kHz wide centered... the general radiated emission limits in § 15.209. (d) A custom built telemetry intentional radiator...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22494738-compositional-dependence-band-gap-ga-nasp-quantum-well-heterostructures','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22494738-compositional-dependence-band-gap-ga-nasp-quantum-well-heterostructures"><span>Compositional dependence of the <span class="hlt">band</span> gap in Ga(NAsP) quantum well heterostructures</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Jandieri, K., E-mail: kakhaber.jandieri@physik.uni-marburg.de; Ludewig, P.; Wegele, T.</p> <p></p> <p>We present experimental and theoretical studies of the composition dependence of the direct <span class="hlt">band</span> gap <span class="hlt">energy</span> in Ga(NAsP)/GaP quantum well heterostructures grown on either (001) GaP- or Si-substrates. The theoretical description takes into account the <span class="hlt">band</span> anti-crossing model for the conduction <span class="hlt">band</span> as well as the modification of the valence subband structure due to the strain resulting from the pseudomorphic epitaxial growth on the respective substrate. The composition dependence of the direct <span class="hlt">band</span> gap of Ga(NAsP) is obtained for a wide range of nitrogen and phosphorus contents relevant for laser applications on Si-substrate.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li class="active"><span>22</span></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li><a href="#" onclick='return showDiv("page_24");'>24</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_22 --> <div id="page_23" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li class="active"><span>23</span></li> <li><a href="#" onclick='return showDiv("page_24");'>24</a></li> <li><a href="#" onclick='return showDiv("page_25");'>25</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="441"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011JChPh.135h5101U','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011JChPh.135h5101U"><span>Pairwise <span class="hlt">additivity</span> of <span class="hlt">energy</span> components in protein-ligand binding: The HIV II protease-Indinavir case</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ucisik, Melek N.; Dashti, Danial S.; Faver, John C.; Merz, Kenneth M.</p> <p>2011-08-01</p> <p>An <span class="hlt">energy</span> expansion (binding <span class="hlt">energy</span> decomposition into n-body interaction terms for n ≥ 2) to express the receptor-ligand binding <span class="hlt">energy</span> for the fragmented HIV II protease-Indinavir system is described to address the role of cooperativity in ligand binding. The outcome of this <span class="hlt">energy</span> expansion is compared to the total receptor-ligand binding <span class="hlt">energy</span> at the Hartree-Fock, density functional theory, and semiempirical levels of theory. We find that the sum of the pairwise interaction <span class="hlt">energies</span> approximates the total binding <span class="hlt">energy</span> to ˜82% for HF and to >95% for both the M06-L density functional and PM6-DH2 semiempirical method. The contribution of the three-body interactions amounts to 18.7%, 3.8%, and 1.4% for HF, M06-L, and PM6-DH2, respectively. We find that the expansion can be safely truncated after n = 3. That is, the contribution of the interactions involving more than three parties to the total binding <span class="hlt">energy</span> of Indinavir to the HIV II protease receptor is negligible. Overall, we find that the two-body terms represent a good approximation to the total binding <span class="hlt">energy</span> of the system, which points to pairwise <span class="hlt">additivity</span> in the present case. This basic principle of pairwise <span class="hlt">additivity</span> is utilized in fragment-based drug design approaches and our results support its continued use. The present results can also aid in the validation of non-bonded terms contained within common force fields and in the correction of systematic errors in physics-based score functions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/12636624','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/12636624"><span><span class="hlt">Energy</span> exchange properties during second-harmonic generation in finite one-dimensional photonic <span class="hlt">band</span>-gap structures with deep gratings.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>D'Aguanno, Giuseppe; Centini, Marco; Scalora, Michael; Sibilia, Concita; Bertolotti, Mario; Bloemer, Mark J; Bowden, Charles M</p> <p>2003-01-01</p> <p>We study second-harmonic generation in finite, one-dimensional, photonic <span class="hlt">band</span>-gap structures with large index contrast in the regime of pump depletion and global phase-matching conditions. We report a number of surprising results: above a certain input intensity, field dynamics resemble a multiwave mixing process, where backward and forward components compete for the available <span class="hlt">energy</span>; the pump field is mostly reflected, revealing a type of optical limiting behavior; and second-harmonic generation becomes balanced in both directions, showing unusual saturation effects with increasing pump intensity. This dynamics was unexpected, and it is bound to influence the way one goes about thinking and designing nonlinear frequency conversion devices in a practical way.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26881883','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26881883"><span>Activation <span class="hlt">energy</span> associated with the electromigration of oligosaccharides through viscosity modifier and polymeric <span class="hlt">additive</span> containing background electrolytes.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kerékgyártó, Márta; Járvás, Gábor; Novák, Levente; Guttman, András</p> <p>2016-02-01</p> <p>The activation <span class="hlt">energy</span> related to the electromigration of oligosaccharides can be determined from their measured electrophoretic mobilities at different temperatures. The effects of a viscosity modifier (ethylene glycol) and a polymeric <span class="hlt">additive</span> (linear polyacrylamide) on the electrophoretic mobility of linear sugar oligomers with α1-4 linked glucose units (maltooligosaccharides) were studied in CE using the activation <span class="hlt">energy</span> concept. The electrophoretic separations of 8-aminopyrene-1,3,6-trisulfonate-labeled maltooligosaccharides were monitored by LIF detection in the temperature range of 20-50°C, using either 0-60% ethylene glycol (viscosity modifier) or 0-3% linear polyacrylamide (polymeric <span class="hlt">additive</span>) containing BGEs. Activation <span class="hlt">energy</span> curves were constructed based on the slopes of the Arrhenius plots. With the use of linear polyacrylamide <span class="hlt">additive</span>, solute size-dependent activation <span class="hlt">energy</span> variations were found for the maltooligosaccharides with polymerization degrees below and above maltoheptaose (DP 7), probably due to molecular conformation changes and possible matrix interaction effects. © 2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28447450','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28447450"><span><span class="hlt">Band</span>-Bending of Ga-Polar GaN Interfaced with Al2O3 through Ultraviolet/Ozone Treatment.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kim, Kwangeun; Ryu, Jae Ha; Kim, Jisoo; Cho, Sang June; Liu, Dong; Park, Jeongpil; Lee, In-Kyu; Moody, Baxter; Zhou, Weidong; Albrecht, John; Ma, Zhenqiang</p> <p>2017-05-24</p> <p>Understanding the <span class="hlt">band</span> bending at the interface of GaN/dielectric under different surface treatment conditions is critically important for device design, device performance, and device reliability. The effects of ultraviolet/ozone (UV/O 3 ) treatment of the GaN surface on the <span class="hlt">energy</span> <span class="hlt">band</span> bending of atomic-layer-deposition (ALD) Al 2 O 3 coated Ga-polar GaN were studied. The UV/O 3 treatment and post-ALD anneal can be used to effectively vary the <span class="hlt">band</span> bending, the valence <span class="hlt">band</span> offset, conduction <span class="hlt">band</span> offset, and the interface dipole at the Al 2 O 3 /GaN interfaces. The UV/O 3 treatment increases the surface <span class="hlt">energy</span> of the Ga-polar GaN, improves the uniformity of Al 2 O 3 deposition, and changes the amount of trapped charges in the ALD layer. The positively charged surface states formed by the UV/O 3 treatment-induced surface factors externally screen the effect of polarization charges in the GaN, in effect, determining the eventual <span class="hlt">energy</span> <span class="hlt">band</span> bending at the Al 2 O 3 /GaN interfaces. An optimal UV/O 3 treatment condition also exists for realizing the "best" interface conditions. The study of UV/O 3 treatment effect on the <span class="hlt">band</span> alignments at the dielectric/III-nitride interfaces will be valuable for applications of transistors, light-emitting diodes, and photovoltaics.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014PhRvL.113w7001N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014PhRvL.113w7001N"><span>Reconstruction of <span class="hlt">Band</span> Structure Induced by Electronic Nematicity in an FeSe Superconductor</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Nakayama, K.; Miyata, Y.; Phan, G. N.; Sato, T.; Tanabe, Y.; Urata, T.; Tanigaki, K.; Takahashi, T.</p> <p>2014-12-01</p> <p>We have performed high-resolution angle-resolved photoemission spectroscopy on an FeSe superconductor (Tc˜8 K ), which exhibits a tetragonal-to-orthorhombic structural transition at Ts˜90 K . At low temperature, we found splitting of the <span class="hlt">energy</span> <span class="hlt">bands</span> as large as 50 meV at the M point in the Brillouin zone, likely caused by the formation of electronically driven nematic states. This <span class="hlt">band</span> splitting persists up to T ˜110 K , slightly above Ts, suggesting that the structural transition is triggered by the electronic nematicity. We have also revealed that at low temperature the <span class="hlt">band</span> splitting gives rise to a van Hove singularity within 5 meV of the Fermi <span class="hlt">energy</span>. The present result strongly suggests that this unusual electronic state is responsible for the unconventional superconductivity in FeSe.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27731531','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27731531"><span>Toward a High-Efficient Utilization of Solar Radiation by Quad-<span class="hlt">Band</span> Solar Spectral Splitting.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Cao, Feng; Huang, Yi; Tang, Lu; Sun, Tianyi; Boriskina, Svetlana V; Chen, Gang; Ren, Zhifeng</p> <p>2016-12-01</p> <p>The promising quad-<span class="hlt">band</span> solar spectral splitter incorporates the properties of the optical filter and the spectrally selective solar thermal absorber can direct PV <span class="hlt">band</span> to PV modules and absorb thermal <span class="hlt">band</span> <span class="hlt">energy</span> for thermal process with low thermal losses. It provides a new strategy for spectral splitting and offers potential ways for hybrid PVT system design. © 2016 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.dtic.mil/docs/citations/ADA412962','DTIC-ST'); return false;" href="http://www.dtic.mil/docs/citations/ADA412962"><span>C (G)-<span class="hlt">Band</span> & X (I) - <span class="hlt">Band</span> Noncoherent Radar Transponder Performance Specification Standard</span></a></p> <p><a target="_blank" href="http://www.dtic.mil/">DTIC Science & Technology</a></p> <p></p> <p>2002-04-01</p> <p>TRAINING RANGE NEVADA TEST SITE STANDARD 262-02 ELECTRONIC TRAJECTORY MEASUREMENTS GROUP C (G) – <span class="hlt">BAND</span> & X (I) – <span class="hlt">BAND</span> NONCOHERENT RADAR...Date 00 Apr 2002 Report Type N/A Dates Covered (from... to) - Title and Subtitle C (G)-<span class="hlt">Band</span> & X (I) - <span class="hlt">Band</span> Noncoherent Radar Transponder...Number of Pages 157 i STANDARD 262-02 C (G) – <span class="hlt">BAND</span> & X (I) – <span class="hlt">BAND</span> NONCOHERENT RADAR TRANSPONDER PERFORMANCE SPECIFICATION STANDARD APRIL 2002 Prepared by</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22494991-sub-band-gap-photo-enhanced-secondary-electron-emission-from-high-purity-single-crystal-chemical-vapor-deposited-diamond','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22494991-sub-band-gap-photo-enhanced-secondary-electron-emission-from-high-purity-single-crystal-chemical-vapor-deposited-diamond"><span>Sub-<span class="hlt">band</span> gap photo-enhanced secondary electron emission from high-purity single-crystal chemical-vapor-deposited diamond</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Yater, J. E., E-mail: joan.yater@nrl.navy.mil; Shaw, J. L.; Pate, B. B.</p> <p>2016-02-07</p> <p>Secondary-electron-emission (SEE) current measured from high-purity, single-crystal (100) chemical-vapor-deposited diamond is found to increase when sub-<span class="hlt">band</span> gap (3.06 eV) photons are incident on the hydrogenated surface. Although the light does not produce photoemission directly, the SEE current increases by more than a factor of 2 before saturating with increasing laser power. In <span class="hlt">energy</span> distribution curves (EDCs), the emission peak shows a corresponding increase in intensity with increasing laser power. However, the emission-onset <span class="hlt">energy</span> in the EDCs remains constant, indicating that the <span class="hlt">bands</span> are pinned at the surface. On the other hand, changes are observed on the high-<span class="hlt">energy</span> side of the distributionmore » as the laser power increases, with a well-defined shoulder becoming more pronounced. From an analysis of this feature in the EDCs, it is deduced that upward <span class="hlt">band</span> bending is present in the near-surface region during the SEE measurements and this <span class="hlt">band</span> bending suppresses the SEE yield. However, sub-<span class="hlt">band</span> gap photon illumination reduces the <span class="hlt">band</span> bending and thereby increases the SEE current. Because the <span class="hlt">bands</span> are pinned at the surface, we conclude that the changes in the <span class="hlt">band</span> levels occur below the surface in the electron transport region. Sample heating produces similar effects as observed with sub-<span class="hlt">band</span> gap photon illumination, namely, an increase in SEE current and a reduction in <span class="hlt">band</span> bending. However, the upward <span class="hlt">band</span> bending is not fully removed by either increasing laser power or temperature, and a minimum <span class="hlt">band</span> bending of ∼0.8 eV is established in both cases. The sub-<span class="hlt">band</span> gap photo-excitation mechanism is under further investigation, although it appears likely at present that defect or gap states play a role in the photo-enhanced SEE process. In the meantime, the study demonstrates the ability of visible light to modify the electronic properties of diamond and enhance the emission capabilities, which may have potential impact for diamond-based vacuum</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014Nanos...6.6531W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014Nanos...6.6531W"><span>2D XANES-XEOL mapping: observation of enhanced <span class="hlt">band</span> gap emission from ZnO nanowire arrays</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wang, Zhiqiang; Guo, Xiaoxuan; Sham, Tsun-Kong</p> <p>2014-05-01</p> <p>Using 2D XANES-XEOL spectroscopy, it is found that the <span class="hlt">band</span> gap emission of ZnO nanowire arrays is substantially enhanced i.e. that the intensity ratio between the <span class="hlt">band</span> gap and defect emissions increases by more than an order of magnitude when the excitation <span class="hlt">energy</span> is scanned across the O K-edge. Possible mechanisms are discussed.Using 2D XANES-XEOL spectroscopy, it is found that the <span class="hlt">band</span> gap emission of ZnO nanowire arrays is substantially enhanced i.e. that the intensity ratio between the <span class="hlt">band</span> gap and defect emissions increases by more than an order of magnitude when the excitation <span class="hlt">energy</span> is scanned across the O K-edge. Possible mechanisms are discussed. Electronic supplementary information (ESI) available: XEOL spectra with different excitation <span class="hlt">energies</span>. X-ray attenuation length vs. photon <span class="hlt">energy</span>. Details of surface defects in ZnO NWs. The second O K-edge and Zn L-edge 2D XANES-XEOL maps. Comparison of the first and second TEY at O K-edge and Zn L-edge scans, respectively. Raman spectra of the ZnO NWs with different IBGE/IDE ratios. See DOI: 10.1039/c4nr01049c</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29789634','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29789634"><span>Giant photovoltaic response in <span class="hlt">band</span> engineered ferroelectric perovskite.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Pal, Subhajit; Swain, Atal Bihari; Biswas, Pranab Parimal; Murali, D; Pal, Arnab; Nanda, B Ranjit K; Murugavel, Pattukkannu</p> <p>2018-05-22</p> <p>Recently the solar <span class="hlt">energy</span>, an inevitable part of green <span class="hlt">energy</span> source, has become a mandatory topics in frontier research areas. In this respect, non-centrosymmetric ferroelectric perovskites with open circuit voltage (V OC ) higher than the bandgap, gain tremendous importance as next generation photovoltaic materials. Here a non-toxic co-doped Ba 1-x (Bi 0.5 Li 0.5 ) x TiO 3 ferroelectric system is designed where the dopants influence the <span class="hlt">band</span> topology in order to enhance the photovoltaic effect. In particular, at the optimal doping concentration (x opt  ~ 0.125) the sample reveals a remarkably high photogenerated field E OC  = 320 V/cm (V OC  = 16 V), highest ever reported in any bulk polycrystalline non-centrosymmetric systems. The <span class="hlt">band</span> structure, examined through DFT calculations, suggests that the shift current mechanism is key to explain the large enhancement in photovoltaic effect in this family.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/18026540','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/18026540"><span>Optical depth measurements by shadow-<span class="hlt">band</span> radiometers and their uncertainties.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Alexandrov, Mikhail D; Kiedron, Peter; Michalsky, Joseph J; Hodges, Gary; Flynn, Connor J; Lacis, Andrew A</p> <p>2007-11-20</p> <p>Shadow-<span class="hlt">band</span> radiometers in general, and especially the Multi-Filter Rotating Shadow-<span class="hlt">band</span> Radiometer (MFRSR), are widely used for atmospheric optical depth measurements. The major programs running MFRSR networks in the United States include the Department of <span class="hlt">Energy</span> Atmospheric Radiation Measurement (ARM) Program, U.S. Department of Agriculture UV-B Monitoring and Research Program, National Oceanic and Atmospheric Administration Surface Radiation (SURFRAD) Network, and NASA Solar Irradiance Research Network (SIRN). We discuss a number of technical issues specific to shadow-<span class="hlt">band</span> radiometers and their impact on the optical depth measurements. These problems include instrument tilt and misalignment, as well as some data processing artifacts. Techniques for data evaluation and automatic detection of some of these problems are described.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018GeoRL..45.1271Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018GeoRL..45.1271Y"><span>Excitation of O+ <span class="hlt">Band</span> EMIC Waves Through H+ Ring Velocity Distributions: Van Allen Probe Observations</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yu, Xiongdong; Yuan, Zhigang; Huang, Shiyong; Yao, Fei; Wang, Dedong; Funsten, Herbert O.; Wygant, John R.</p> <p>2018-02-01</p> <p>A typical case of electromagnetic ion cyclotron (EMIC) emissions with both He+ <span class="hlt">band</span> and O+ <span class="hlt">band</span> waves was observed by Van Allen Probe A on 14 July 2014. These emissions occurred in the morning sector on the equator inside the plasmasphere, in which region O+ <span class="hlt">band</span> EMIC waves prefer to appear. Through property analysis of these emissions, it is found that the He+ <span class="hlt">band</span> EMIC waves are linearly polarized and propagating quasi-parallelly along the background magnetic field, while the O+ <span class="hlt">band</span> ones are of linear and left-hand polarization and propagating obliquely with respect to the background magnetic field. Using the in situ observations of plasma environment and particle data, excitation of these O+ <span class="hlt">band</span> EMIC waves has been investigated with the linear growth theory. The calculated linear growth rate shows that these O+ <span class="hlt">band</span> EMIC waves can be locally excited by ring current protons with ring velocity distributions. The comparison of the observed wave spectral intensity and the calculated growth rate suggests that the density of H+ rings providing the free <span class="hlt">energy</span> for the instability has decreased after the wave grows. Therefore, this paper provides a direct observational evidence to the excitation mechanism of O+ <span class="hlt">band</span> EMIC waves: ring current protons with ring distributions provide the free <span class="hlt">energy</span> supporting the instability in the presence of rich O+ in the plasmasphere.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19770046217&hterms=kinetic+energy&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3Dkinetic%2Benergy','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19770046217&hterms=kinetic+energy&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3Dkinetic%2Benergy"><span>Conversion of laser <span class="hlt">energy</span> to gas kinetic <span class="hlt">energy</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Caledonia, G. E.</p> <p>1977-01-01</p> <p>Techniques for the gas-phase absorption of laser <span class="hlt">energy</span> with ultimate conversion to heat or directed kinetic <span class="hlt">energy</span> are reviewed. It is shown that the efficiency of resonance absorption by the vibration/rotation <span class="hlt">bands</span> of the working gas can be enhanced by operating at sufficiently high pressures so that the linewidths of the absorbing transition exceed the line spacing. Within this limit, the gas can absorb continuously over the full spectral region of the <span class="hlt">band</span>, and bleaching can be minimized since the manifold of molecular vibrational levels can simultaneously absorb the laser radiation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/9268442','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/9268442"><span>Coloured leg <span class="hlt">bands</span> affect male mate-guarding behaviour in the bluethroat</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Johnsen; Lifjeld; Rohde</p> <p>1997-07-01</p> <p>Artificial traits such as coloured leg <span class="hlt">bands</span> may affect an individual's mating success, as shown for some birds. One explanation is that colour-matching with a sexual ornament affects the individual's sexual attractiveness. This study reports a colour-<span class="hlt">band</span> experiment with free-living bluethroats, Luscinia s. svecicaa species where males have a distinct blue and chestnut throat and upper breast. There was no apparent difference in pairing success between males with ornament-matching colour <span class="hlt">bands</span> (blue and orange) and males with non-ornamental colour <span class="hlt">bands</span>. However, males with ornamental <span class="hlt">bands</span> guarded their mates less intensely and spent more time singing, performing song flights and intruding into neighbours' territories than males with non-ornamental <span class="hlt">bands</span>. We conclude that colour <span class="hlt">bands</span> affect the trade-off between mate guarding and advertisement behaviour in a way that is consistent with the hypothesis that <span class="hlt">bands</span> with ornamental colours improve a male's attractiveness. The results are in concordance with a previous study of the same population, showing that males with experimentally reduced attractiveness guarded their mates more closely and advertised less for <span class="hlt">additional</span> mates, than non-manipulated males.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/1211387-ultrafast-band-gap-oscillations-iron-pyrite','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/1211387-ultrafast-band-gap-oscillations-iron-pyrite"><span>Ultrafast <span class="hlt">band</span>-gap oscillations in iron pyrite</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Kolb, B; Kolpak, AM</p> <p>2013-12-20</p> <p>With its combination of favorable <span class="hlt">band</span> gap, high absorption coefficient, material abundance, and low cost, iron pyrite, FeS2, has received a great deal of attention over the past decades as a promising material for photovoltaic applications such as solar cells and photoelectrochemical cells. Devices made from pyrite, however, exhibit open circuit voltages significantly lower than predicted, and despite a recent resurgence of interest in the material, there currently exists no widely accepted explanation for this disappointing behavior. In this paper, we show that phonons, which have been largely overlooked in previous efforts, may play a significant role. Using fully self-consistentmore » GW calculations, we demonstrate that a phonon mode related to the oscillation of the sulfur-sulfur bond distance in the pyrite structure is strongly coupled to the <span class="hlt">energy</span> of the conduction-<span class="hlt">band</span> minimum, leading to an ultrafast (approximate to 100 fs) oscillation in the <span class="hlt">band</span> gap. Depending on the coherency of the phonons, we predict that this effect can cause changes of up to +/- 0.3 eV relative to the accepted FeS2 <span class="hlt">band</span> gap at room temperature. Harnessing this effect via temperature or irradiation with infrared light could open up numerous possibilities for novel devices such as ultrafast switches and adaptive solar absorbers.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018ISPAr42.3.1085L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018ISPAr42.3.1085L"><span>Sharpending of the Vnir and SWIR <span class="hlt">Bands</span> of the Wide <span class="hlt">Band</span> Spectral Imager Onboard Tiangong-II Imagery Using the Selected <span class="hlt">Bands</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Liu, Q.; Li, X.; Liu, G.; Huang, C.; Li, H.; Guan, X.</p> <p>2018-04-01</p> <p>The Tiangong-II space lab was launched at the Jiuquan Satellite Launch Center of China on September 15, 2016. The Wide <span class="hlt">Band</span> Spectral Imager (WBSI) onboard the Tiangong-II has 14 visible and near-infrared (VNIR) spectral <span class="hlt">bands</span> covering the range from 403-990 nm and two shortwave infrared (SWIR) <span class="hlt">bands</span> covering the range from 1230-1250 nm and 1628-1652 nm respectively. In this paper the selected <span class="hlt">bands</span> are proposed which aims at considering the closest spectral similarities between the VNIR with 100 m spatial resolution and SWIR <span class="hlt">bands</span> with 200 m spatial resolution. The evaluation of Gram-Schmidt transform (GS) sharpening techniques embedded in ENVI software is presented based on four types of the different low resolution pan <span class="hlt">band</span>. The experimental results indicated that the VNIR <span class="hlt">band</span> with higher CC value with the raw SWIR <span class="hlt">Band</span> was selected, more texture information was injected the corresponding sharpened SWIR <span class="hlt">band</span> image, and at that time another sharpened SWIR <span class="hlt">band</span> image preserve the similar spectral and texture characteristics to the raw SWIR <span class="hlt">band</span> image.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1996ApSS..104..595D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1996ApSS..104..595D"><span>Low-<span class="hlt">energy</span> yield spectroscopy determination of <span class="hlt">band</span> offsets: application to the epitaxial Ge/Si(100) heterostructure</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Di Gaspare, L.; Capellini, G.; Chudoba, C.; Sebastiani, M.; Evangelisti, F.</p> <p>1996-09-01</p> <p>We apply a new experimental method for determining <span class="hlt">band</span> lineups at the Ge/Si(100) heterostructure. This method uses a modern version of an old spectroscopy: the photoelectric yield spectroscopy excited with photons in the near UV range. It is shown that both substrate and overlayer valence-<span class="hlt">band</span> tops can be identified in the yield spectrum, thus allowing a direct and precise determination of the <span class="hlt">band</span> lineup. We find an offset of 0.36 ± 0.02 eV for heterojunctions whose overlayers were grown according to the Stranski-Krastanov mechanism.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27129802','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27129802"><span>Management of Gastric Obstruction Caused by Adjustable Gastric <span class="hlt">Band</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Czeiger, David; Abu-Swis, Shadi; Shaked, Gad; Ovnat, Amnon; Sebbag, Gilbert</p> <p>2016-12-01</p> <p>Optimal adjustment of the filling volume of laparoscopic adjustable gastric <span class="hlt">banding</span> is challenging and commonly performed empirically. Patients with <span class="hlt">band</span> over-inflation and gastric obstruction arrive at the emergency department complaining of recurrent vomiting. In cases of gastric obstruction, intra-<span class="hlt">band</span> pressure measurement may assist in determining the amount of fluid that should be removed from the <span class="hlt">band</span>; however, our investigations have determined that intra-<span class="hlt">band</span> pressure assessment need not play a role in the treatment of gastric <span class="hlt">band</span> obstruction. In patients coming to the emergency department with gastric <span class="hlt">band</span> obstruction, we measured intra-<span class="hlt">band</span> pressure at arrival and following stepped removal of fluid, comparing the initial pressure with post-deflation pressure and measuring the volume of fluid removed. Forty-eight patients participated in the study. Forty-five patients had a low-pressure/high-volume <span class="hlt">band</span>. Their mean baseline pressure was 54.6 ± 22.3 mmHg. The mean volume of fluid removed from the <span class="hlt">band</span> was 1.3 ± 0.8 ml. The mean post-deflation pressure was 22.5 ± 16.3 mmHg. Nearly 30 % of patients required as little as 0.5 ml of fluid removal, and 60 % of them were free of symptoms with removal of 1 ml. Our results indicate that intra-<span class="hlt">band</span> pressure measurement is of little value for determining the amount of fluid that should be removed for treatment of <span class="hlt">band</span> obstruction. We suggest the removal of fluid in volumes of 0.5 ml until symptoms are relieved. Only in complicated cases, such as in patients having recurrent obstructions, should <span class="hlt">additional</span> modalities be employed for further management guidance.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JSV...413..101W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JSV...413..101W"><span>Effect of thermal stresses on frequency <span class="hlt">band</span> structures of elastic metamaterial plates</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wu, Ying; Yu, Kaiping; Yang, Linyun; Zhao, Rui; Shi, Xiaotian; Tian, Kuo</p> <p>2018-01-01</p> <p>We investigate the effect of thermal stresses on the <span class="hlt">band</span> structure of elastic metamaterial plates by developing a useful finite-element based method. The thermal field is assumed to be uniform throughout the whole plate. Specifically, we find that the stiffness matrix of plate element is comprised of elastic and thermal stresses parts, which can be regarded as a linear function of temperature difference. We <span class="hlt">additionally</span> demonstrate that the relative magnitudes between elastic properties and thermal stresses will lead to nonlinear effects on frequency <span class="hlt">band</span> structures based on two different types of metamaterial plates made of single and double inclusions of square plates, respectively. Then, we validate the proposed approach by comparing the <span class="hlt">band</span> structures with the frequency response curves obtained in finite periodic structures. We conduct sensitivity analysis and discuss in-depth the sensitivities of <span class="hlt">band</span> structures with respect to temperature difference to quantitatively investigate the effect of thermal stresses on each <span class="hlt">band</span>. In <span class="hlt">addition</span>, the coupled effects of thermal stresses and temperature-dependent material properties on the <span class="hlt">band</span> structure of Aluminum/silicone rubber plate have also been discussed. The proposed method and new findings in this paper extends the ability of existing metamaterial plates by enabling tunability over a wide range of frequencies in thermal environments.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/859220','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/859220"><span>Renewable <span class="hlt">Energy</span> Development on Tribal Lamds of Viejas</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Terrence Meyer; Mike Elenbaas</p> <p>2005-09-30</p> <p>The purpose of this study is to investigate the feasibility of Renewable <span class="hlt">Energy</span> Development on the lands of the Viejas <span class="hlt">Band</span> of the Kumeyaay Indian Nation. In <span class="hlt">addition</span>, the study will investigate the feasibility of forming a renewable <span class="hlt">energy</span> based tribal utility. Viejas contracted with Black & Veatch and Fredericks, Pelcyger & Hester, LLC to assist in the development of a feasibility study to ascertain the economics and operational factors of forming an electric and water utility. This report is the result of the investigation conducted by Black & Veatch, with input from Viejas Tribal Government.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li class="active"><span>23</span></li> <li><a href="#" onclick='return showDiv("page_24");'>24</a></li> <li><a href="#" onclick='return showDiv("page_25");'>25</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_23 --> <div id="page_24" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li class="active"><span>24</span></li> <li><a href="#" onclick='return showDiv("page_25");'>25</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="461"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/863216','DOE-PATENT-XML'); return false;" href="https://www.osti.gov/servlets/purl/863216"><span>Inertial <span class="hlt">energy</span> storage device</span></a></p> <p><a target="_blank" href="http://www.osti.gov/doepatents">DOEpatents</a></p> <p>Knight, Jr., Charles E.; Kelly, James J.; Pollard, Roy E.</p> <p>1978-01-01</p> <p>The inertial <span class="hlt">energy</span> storage device of the present invention comprises a composite ring formed of circumferentially wound resin-impregnated filament material, a flanged hollow metal hub concentrically disposed in the ring, and a plurality of discrete filament bandsets coupling the hub to the ring. Each bandset is formed of a pair of parallel <span class="hlt">bands</span> affixed to the hub in a spaced apart relationship with the axis of rotation of the hub being disposed between the <span class="hlt">bands</span> and with each <span class="hlt">band</span> being in the configuration of a hoop extending about the ring along a chordal plane thereof. The bandsets are disposed in an angular relationship with one another so as to encircle the ring at spaced-apart circumferential locations while being disposed in an overlapping relationship on the flanges of the hub. The <span class="hlt">energy</span> storage device of the present invention has the capability of substantial <span class="hlt">energy</span> storage due to the relationship of the filament <span class="hlt">bands</span> to the ring and the flanged hub.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/5224838','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/5224838"><span>Temporal patterns of apparent leg <span class="hlt">band</span> retention in North American geese</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Zimmerman, Guthrie S.; Kendall, William L.; Moser, Timothy J.; White, Gary C.; Doherty, Paul F.</p> <p>2009-01-01</p> <p>An important assumption of mark?recapture studies is that individuals retain their marks, which has not been assessed for goose reward <span class="hlt">bands</span>. We estimated aluminum leg <span class="hlt">band</span> retention probabilities and modeled how <span class="hlt">band</span> retention varied with <span class="hlt">band</span> type (standard vs. reward <span class="hlt">band</span>), <span class="hlt">band</span> age (1-40 months), and goose characteristics (species and size class) for Canada (Branta canadensis), cackling (Branta hutchinsii), snow (Chen caerulescens), and Ross?s (Chen rossii) geese that field coordinators double-leg <span class="hlt">banded</span> during a North American goose reward <span class="hlt">band</span> study (N = 40,999 individuals from 15 populations). We conditioned all models in this analysis on geese that were encountered with >1 leg <span class="hlt">band</span> still attached (n = 5,747 dead recoveries and live recaptures). Retention probabilities for standard aluminum leg <span class="hlt">bands</span> were high (estimate of 0.9995, SE = 0.001) and constant over 1-40 months. In contrast, apparent retention probabilities for reward <span class="hlt">bands</span> demonstrated an interactive relationship between 5 size and species classes (small cackling, medium Canada, large Canada, snow, and Ross?s geese). In <span class="hlt">addition</span>, apparent retention probabilities for each of the 5 classes varied quadratically with time, being lower immediately after <span class="hlt">banding</span> and at older age classes. The differential retention probabilities among <span class="hlt">band</span> type (reward vs. standard) that we observed suggests that 1) models estimating reporting probability should incorporate differential <span class="hlt">band</span> loss if it is nontrivial, 2) goose managers should consider the costs and benefits of double-<span class="hlt">banding</span> geese on an operational basis, and 3) the United States Geological Survey Bird <span class="hlt">Banding</span> Lab should modify protocols for receiving recovery data.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1990AIPC..200...36W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1990AIPC..200...36W"><span>Nature of the valence <span class="hlt">band</span> states in Bi2(Ca, Sr, La)3Cu2O8</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wells, B. O.; Lindberg, P. A. P.; Shen, Z.-X.; Dessau, D. S.; Spicer, W. E.; Lindau, I.; Mitzi, D. B.; Kapitulnik, A.</p> <p>1990-01-01</p> <p>We have used photoemission spectroscopy to examine the symmetry of the occupied states of the valence <span class="hlt">band</span> for the La doped superconductor Bi2(Ca, Sr, La)3Cu2O8. While the oxygen states near the bottom of the 7 eV wide valence <span class="hlt">band</span> exhibit predominantly O 2pz symmetry, the states at the top of the valence <span class="hlt">band</span> extending to the Fermi level are found to have primarily O 2px and O 2py character. We have also examined anomalous intensity enhancements in the valence <span class="hlt">band</span> feature for photon <span class="hlt">energies</span> near 18 eV. These enhancements, which occur at photon <span class="hlt">energies</span> ranging from 15.8 to 18.0 eV for the different valence <span class="hlt">band</span> features, are not consistent with either simple final state effects or direct O2s transitions to unoccupied O2p states.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012ChPhB..21e4101Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012ChPhB..21e4101Z"><span>Dual-<span class="hlt">band</span> frequency selective surface with large <span class="hlt">band</span> separation and stable performance</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zhou, Hang; Qu, Shao-Bo; Peng, Wei-Dong; Lin, Bao-Qin; Wang, Jia-Fu; Ma, Hua; Zhang, Jie-Qiu; Bai, Peng; Wang, Xu-Hua; Xu, Zhuo</p> <p>2012-05-01</p> <p>A new technique of designing a dual-<span class="hlt">band</span> frequency selective surface with large <span class="hlt">band</span> separation is presented. This technique is based on a delicately designed topology of L- and Ku-<span class="hlt">band</span> microwave filters. The two <span class="hlt">band</span>-pass responses are generated by a capacitively-loaded square-loop frequency selective surface and an aperture-coupled frequency selective surface, respectively. A Faraday cage is located between the two frequency selective surface structures to eliminate undesired couplings. Based on this technique, a dual-<span class="hlt">band</span> frequency selective surface with large <span class="hlt">band</span> separation is designed, which possesses large <span class="hlt">band</span> separation, high selectivity, and stable performance under various incident angles and different polarizations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/21828568','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21828568"><span>A simple <span class="hlt">energy</span> filter for low <span class="hlt">energy</span> electron microscopy/photoelectron emission microscopy instruments.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Tromp, R M; Fujikawa, Y; Hannon, J B; Ellis, A W; Berghaus, A; Schaff, O</p> <p>2009-08-05</p> <p><span class="hlt">Addition</span> of an electron <span class="hlt">energy</span> filter to low <span class="hlt">energy</span> electron microscopy (LEEM) and photoelectron emission microscopy (PEEM) instruments greatly improves their analytical capabilities. However, such filters tend to be quite complex, both electron optically and mechanically. Here we describe a simple <span class="hlt">energy</span> filter for the existing IBM LEEM/PEEM instrument, which is realized by adding a single scanning aperture slit to the objective transfer optics, without any further modifications to the microscope. This <span class="hlt">energy</span> filter displays a very high <span class="hlt">energy</span> resolution ΔE/E = 2 × 10(-5), and a non-isochromaticity of ∼0.5 eV/10 µm. The setup is capable of recording selected area electron <span class="hlt">energy</span> spectra and angular distributions at 0.15 eV <span class="hlt">energy</span> resolution, as well as <span class="hlt">energy</span> filtered images with a 1.5 eV <span class="hlt">energy</span> pass <span class="hlt">band</span> at an estimated spatial resolution of ∼10 nm. We demonstrate the use of this <span class="hlt">energy</span> filter in imaging and spectroscopy of surfaces using a laboratory-based He I (21.2 eV) light source, as well as imaging of Ag nanowires on Si(001) using the 4 eV <span class="hlt">energy</span> loss Ag plasmon.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/6817137-quasiparticle-band-offset-interface-band-gaps-ultrathin-superlattices-gaas-alas-heterojunctions','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/6817137-quasiparticle-band-offset-interface-band-gaps-ultrathin-superlattices-gaas-alas-heterojunctions"><span>Quasiparticle <span class="hlt">band</span> offset at the (001) interface and <span class="hlt">band</span> gaps in ultrathin superlattices of GaAs-AlAs heterojunctions</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Zhang, S.B.; Cohen, M.L.; Louie, S.G.</p> <p>1990-05-15</p> <p>A newly developed first-principles quasiparticle theory is used to calculate the <span class="hlt">band</span> offset at the (001) interface and <span class="hlt">band</span> gaps in 1{times}1 and 2{times}2 superlattices of GaAs-AlAs heterojunctions. We find a sizable many-body contribution to the valence-<span class="hlt">band</span> offset which is dominated by the many-body corrections to bulk GaAs and AlAs quasiparticle <span class="hlt">energies</span>. The resultant offset {Delta}{ital E}{sub {ital v}}=0.53{plus minus}0.05 eV is in good agreement with the recent experimental values of 0.50--0.56 eV. Our calculated direct <span class="hlt">band</span> gaps for ultrathin superlattices are also in good agreement with experiment. The {ital X}{sub 1{ital c}}-derived state at point {bar {Gamma}}, is however,more » above the {Gamma}{sub 1{ital c}}-derived state for both the 1{times}1 and 2{times}2 lattices, contrary to results obtained under the virtual-crystal approximation (a limiting case for the Kronig-Penny model) and some previous local-density-approximation (corrected) calculations. The differences are explained in terms of atomic-scale localizations and many-body effects. Oscillator strengths and the effects of disorder on the spectra are discussed.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018AIPA....8e5003G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AIPA....8e5003G"><span>Theoretical analysis on lower <span class="hlt">band</span> cascade as a mechanism for multiband chorus in the Earth's magnetosphere</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gao, Xinliang; Lu, Quanming; Wang, Shaojie; Wang, Shui</p> <p>2018-05-01</p> <p>Whistler-mode waves play a crucial role in controlling electron dynamics in the Earth's Van Allen radiation belt, which is increasingly important for spacecraft safety. Using THEMIS waveform data, Gao et al. [X. L. Gao, Q. Lu, J. Bortnik, W. Li, L. Chen, and S. Wang, Geophys. Res. Lett., 43, 2343-2350, 2016] have reported two multiband chorus events, wherein upper-<span class="hlt">band</span> chorus appears at harmonics of lower-<span class="hlt">band</span> chorus. They proposed that upper-<span class="hlt">band</span> harmonic waves are excited through the nonlinear coupling between the electromagnetic and electrostatic components of lower-<span class="hlt">band</span> chorus, a second-order effect called "lower <span class="hlt">band</span> cascade". However, the theoretical explanation of lower <span class="hlt">band</span> cascade was not thoroughly explained in the earlier work. In this paper, based on a cold plasma assumption, we have obtained the explicit nonlinear driven force of lower <span class="hlt">band</span> cascade through a full nonlinear theoretical analysis, which includes both the ponderomotive force and coupling between electrostatic and electromagnetic components of the pump whistler wave. Moreover, we discover the existence of an efficient <span class="hlt">energy</span>-transfer (E-t) channel from lower-<span class="hlt">band</span> to upper-<span class="hlt">band</span> whistler-mode waves during lower <span class="hlt">band</span> cascade for the first time, which is also confirmed by PIC simulations. For lower-<span class="hlt">band</span> whistler-mode waves with a small wave normal angle (WNA), the E-t channel is detected when the driven upper-<span class="hlt">band</span> wave nearly satisfies the linear dispersion relation of whistler mode. While, for lower-<span class="hlt">band</span> waves with a large WNA, the E-t channel is found when the lower-<span class="hlt">band</span> wave is close to its resonant frequency, and the driven upper-<span class="hlt">band</span> wave becomes quasi-electrostatic. Through this efficient channel, the harmonic upper <span class="hlt">band</span> of whistler waves is generated through <span class="hlt">energy</span> cascade from the lower <span class="hlt">band</span>, and the two-<span class="hlt">band</span> spectral structure of whistler waves is then formed. Both two types of <span class="hlt">banded</span> whistler-mode spectrum have also been successfully reproduced by PIC simulations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFMSM31A2476G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFMSM31A2476G"><span>Generation of Multi-<span class="hlt">band</span> Chorus by Lower <span class="hlt">Band</span> Cascade in the Earth's Magnetosphere</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gao, X.; Lu, Q.; Chen, L.; Bortnik, J.; Li, W.; Wang, S.</p> <p>2016-12-01</p> <p>Chorus waves are intense electromagnetic whistler-mode emissions in the magnetosphere, typically falling into two distinct frequency <span class="hlt">bands</span>: a lower <span class="hlt">band</span> (0.1-0.5fce) and an upper <span class="hlt">band</span> (0.5-0.8fce) with a power gap at about 0.5fce. In this letter, with the THEMIS satellite, we observed two special chorus events, which are called as multi-<span class="hlt">band</span> chorus because upper <span class="hlt">band</span> chorus is located at harmonics of lower <span class="hlt">band</span> chorus. We propose a new potential generation mechanism for multi-<span class="hlt">band</span> chorus, which is called as lower <span class="hlt">band</span> cascade. In this scenario, a density mode with a frequency equal to that of lower <span class="hlt">band</span> chorus is caused by the ponderomotive effect (inhomogeneity of the electric amplitude) along the wave vector, and then upper <span class="hlt">band</span> chorus with the frequency twice that of lower <span class="hlt">band</span> chorus is generated through wave-wave couplings between lower <span class="hlt">band</span> chorus and the density mode. The mechanism provides a new insight into the evolution of whistler-mode chorus in the Earth's magnetosphere.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017SuMi..112..328L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017SuMi..112..328L"><span><span class="hlt">Band</span> gap engineering of BC2N for nanoelectronic applications</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lim, Wei Hong; Hamzah, Afiq; Ahmadi, Mohammad Taghi; Ismail, Razali</p> <p>2017-12-01</p> <p>The BC2N as an example of boron-carbon-nitride (BCN), has the analogous structure as the graphene and boron nitride. It is predicted to have controllable electronic properties. Therefore, the analytical study on the engineer-able <span class="hlt">band</span> gap of the BC2N is carried out based on the schematic structure of BC2N. The Nearest Neighbour Tight Binding (NNTB) model is employed with the dispersion relation and the density of state (DOS) as the main <span class="hlt">band</span> gap analysing parameter. The results show that the hopping integrals having the significant effect on the <span class="hlt">band</span> gap, <span class="hlt">band</span> structure and DOS of BC2N nanowire (BC2NNW) need to be taken into consideration. The presented model indicates consistent trends with the published computational results around the Dirac points with the extracted <span class="hlt">band</span> gap of 0.12 eV. Also, it is distinguished that wide <span class="hlt">energy</span> gap of boron nitride (BN) is successfully narrowed by this carbon doped material which assures the application of BC2N on the nanoelectronics and optoelectronics in the near future.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/874723','DOE-PATENT-XML'); return false;" href="https://www.osti.gov/servlets/purl/874723"><span>Permanent magnet focused X-<span class="hlt">band</span> photoinjector</span></a></p> <p><a target="_blank" href="http://www.osti.gov/doepatents">DOEpatents</a></p> <p>Yu, David U. L.; Rosenzweig, James</p> <p>2002-09-10</p> <p>A compact high <span class="hlt">energy</span> photoelectron injector integrates the photocathode directly into a multicell linear accelerator with no drift space between the injection and the linac. High electron beam brightness is achieved by accelerating a tightly focused electron beam in an integrated, multi-cell, X-<span class="hlt">band</span> rf linear accelerator (linac). The photoelectron linac employs a Plane-Wave-Transformer (PWT) design which provides strong cell-to-cell coupling, easing manufacturing tolerances and costs.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1989JMoSp.138...19F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1989JMoSp.138...19F"><span>Near-infrared emission <span class="hlt">bands</span> of TeH and TeD</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Fink, E. H.; Setzer, K. D.; Ramsay, D. A.; Vervloet, M.</p> <p>1989-11-01</p> <p>High-resolution emission spectra of TeH and TeD have been obtained in the region 4200 to 3600 cm -1 using a Bomem DA3.002 Fourier transform spectrometer. Analyses are given for the 0-0 and 1-1 <span class="hlt">bands</span> of the X 22Π{1}/{2}-X 12Π{3}/{2} system of TeH and for the 0-0 <span class="hlt">band</span> of TeD. In <span class="hlt">addition</span> the 2-0 vibrational overtone <span class="hlt">bands</span> of 130TeH, 128TeH, and 126TeH are observed and analyzed. Accurate molecular constants are given for the first time.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20070022570&hterms=fingerprints&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3Dfingerprints','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20070022570&hterms=fingerprints&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3Dfingerprints"><span>Chromosome Aberrations in Human Epithelial Cells Exposed Los Alamos High-<span class="hlt">Energy</span> Secondary Neutrons: M-<span class="hlt">BAND</span> Analysis</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Hada, M.; Saganti, P. B.; Gersey, B.; Wilkins, R.; Cucinotta, F. A.; Wu, H.</p> <p>2007-01-01</p> <p>High-<span class="hlt">energy</span> secondary neutrons, produced by the interaction of galactic cosmic rays (GCR) with the atmosphere, spacecraft structure and planetary surfaces, contribute a significant fraction to the dose equivalent radiation measurement in crew members and passengers of commercial aviation travel as well as astronauts in space missions. The Los Alamos Nuclear Science Center (LANSCE) neutron facility's 30L beam line (4FP30L-A/ICE House) is known to generate neutrons that simulate the secondary neutron spectrum of the Earth's atmosphere at high altitude. The neutron spectrum is also similar to that measured onboard spacecrafts like the MIR and the International Space Station (ISS). To evaluate the biological damage, we exposed human epithelial cells in vitro to the LANSCE neutron beams with an entrance dose rate of 2.5 cGy/hr, and studied the induction of chromosome aberrations that were identified with multicolor-<span class="hlt">banding</span> in situ hybridization (m<span class="hlt">BAND</span>) technique. With this technique, individually painted chromosomal <span class="hlt">bands</span> on one chromosome allowed the identification of inter-chromosomal aberrations (translocation to unpainted chromosomes) and intra-chromosomal aberrations (inversions and deletions within a single painted chromosome). Compared to our previous results with gamma-rays and 600 MeV/nucleon Fe ions of high dose rate at NSRL (NASA Space Radiation Laboratory at Brookhaven National Laboratory), the neutron data from the LANSCE experiments showed significantly higher frequency of chromosome aberrations. However, detailed analysis of the inversion type revealed that all of the three radiation types in the study induced a low incidence of simple inversions. Most of the inversions in gamma-ray irradiated samples were accompanied by other types of intrachromosomal aberrations but few inversions were accompanied by interchromosomal aberrations. In contrast, neutrons and Fe ions induced a significant fraction of inversions that involved complex rearrangements of both</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23573244','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23573244"><span>Increased postprandial <span class="hlt">energy</span> expenditure may explain superior long term weight loss after Roux-en-Y gastric bypass compared to vertical <span class="hlt">banded</span> gastroplasty.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Werling, Malin; Olbers, Torsten; Fändriks, Lars; Bueter, Marco; Lönroth, Hans; Stenlöf, Kaj; le Roux, Carel W</p> <p>2013-01-01</p> <p>Gastric bypass results in greater weight loss than Vertical <span class="hlt">banded</span> gastroplasty (VBG), but the underlying mechanisms remain unclear. In <span class="hlt">addition</span> to effects on <span class="hlt">energy</span> intake the two bariatric techniques may differentially influence <span class="hlt">energy</span> expenditure (EE). Gastric bypass in rats increases postprandial EE enough to result in elevated EE over 24 hours. This study aimed to investigate alterations in postprandial EE after gastric bypass and VBG in humans. Fourteen women from a randomized clinical trial between gastric bypass (n = 7) and VBG (n = 7) were included. Nine years postoperatively and at weight stability patients were assessed for body composition and calorie intake. EE was measured using indirect calorimetry in a respiratory chamber over 24 hours and focused on the periods surrounding meals and sleep. Blood samples were analysed for postprandial gut hormone responses. Groups did not differ regarding body composition or food intake either preoperatively or at study visit. Gastric bypass patients had higher EE postprandially (p = 0.018) and over 24 hours (p = 0.048) compared to VBG patients. Postprandial peptide YY (PYY) and glucagon like peptide 1 (GLP-1) levels were higher after gastric bypass (both p<0.001). Gastric bypass patients have greater meal induced EE and total 24 hours EE compared to VBG patients when assessed 9 years postoperatively. Postprandial satiety gut hormone responses were exaggerated after gastric bypass compared to VBG. Long-term weight loss maintenance may require significant changes in several physiological mechanisms which will be important to understand if non-surgical approaches are to mimic the effects of bariatric surgery.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27490919','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27490919"><span><span class="hlt">Energy</span> Metabolic Adaptation and Cardiometabolic Improvements One Year After Gastric Bypass, Sleeve Gastrectomy, and Gastric <span class="hlt">Band</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Tam, Charmaine S; Redman, Leanne M; Greenway, Frank; LeBlanc, Karl A; Haussmann, Mark G; Ravussin, Eric</p> <p>2016-10-01</p> <p>It is not known whether the magnitude of metabolic adaptation, a greater than expected drop in <span class="hlt">energy</span> expenditure, depends on the type of bariatric surgery and is associated with cardiometabolic improvements. To compare changes in <span class="hlt">energy</span> expenditure (metabolic chamber) and circulating cardiometabolic markers 8 weeks and 1 year after Roux-en-y bypass (RYGB), sleeve gastrectomy (SG), laparoscopic adjustable gastric <span class="hlt">band</span> (LAGB), or a low-calorie diet (LCD). Design, Setting, Participants, and Intervention: This was a parallel-arm, prospective observational study of 30 individuals (27 females; mean age, 46 ± 2 years; body mass index, 47.2 ± 1.5 kg/m 2 ) either self-selecting bariatric surgery (five RYGB, nine SG, seven LAGB) or on a LCD (n = 9) intervention (800 kcal/d for 8 weeks, followed by weight maintenance). After 1 year, the RYGB and SG groups had similar degrees of body weight loss (33-36%), whereas the LAGB and LCD groups had 16 and 4% weight loss, respectively. After adjusting for changes in body composition, 24-hour <span class="hlt">energy</span> expenditure was significantly decreased in all treatment groups at 8 weeks (-254 to -82 kcal/d), a drop that only persisted in RYGB (-124 ± 42 kcal/d; P = .002) and SG (-155 ± 118 kcal/d; P = .02) groups at 1 year. The degree of metabolic adaptation (24-hour and sleeping <span class="hlt">energy</span> expenditure) was not significantly different between the treatment groups at either time-point. Plasma high-density lipoprotein and total and high molecular weight adiponectin were increased, and triglycerides and high-sensitivity C-reactive protein levels were reduced 1 year after RYGB or SG. Metabolic adaptation of approximately 150 kcal/d occurs after RYGB and SG surgery. Future studies are required to examine whether these effects remain beyond 1 year.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29561142','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29561142"><span>Intermediate <span class="hlt">Band</span> Material of Titanium-Doped Tin Disulfide for Wide Spectrum Solar Absorption.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Hu, Keyan; Wang, Dong; Zhao, Wei; Gu, Yuhao; Bu, Kejun; Pan, Jie; Qin, Peng; Zhang, Xian; Huang, Fuqiang</p> <p>2018-04-02</p> <p>Intermediate <span class="hlt">band</span> (IB) materials are of great significance due to their superior solar absorption properties. Here, two IBs peaking at 0.88 and 1.33 eV are reported to be present in the forbidden gap of semiconducting SnS 2 ( E g = 2.21 eV) by doping titanium up to 6 atom % into the Sn site via a solid-state reaction at 923 K. The solid solution of Sn 1- x Ti x S 2 is able to be formed, which is attributed to the isostructural structure of SnS 2 and TiS 2 . These two IBs were detected in the UV-vis-NIR absorption spectra with the appearance of two <span class="hlt">additional</span> absorption responses at the respective regions, which in good agreement with the conclusion of first-principles calculations. The valence <span class="hlt">band</span> maximum (VBM) consists mostly of the S 3p state, and the conduction <span class="hlt">band</span> minimum (CBM) is the hybrid state composing of Ti 3d (e g ), S 3p, and Sn 5s, and the IBs are mainly the nondegenerate t 2g states of Ti 3d orbitals. The electronic states of Ti 3d reveal a good ability to transfer electrons between metal and S atoms. These wide-spectrum absorption IBs bring about more solar <span class="hlt">energy</span> utilization to enhance solar thermal collection and photocatalytic degradation of methyl orange.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/17009489','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/17009489"><span>Utilizing gamma <span class="hlt">band</span> to improve mental task based brain-computer interface design.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Palaniappan, Ramaswamy</p> <p>2006-09-01</p> <p>A common method for designing brain-computer Interface (BCI) is to use electroencephalogram (EEG) signals extracted during mental tasks. In these BCI designs, features from EEG such as power and asymmetry ratios from delta, theta, alpha, and beta <span class="hlt">bands</span> have been used in classifying different mental tasks. In this paper, the performance of the mental task based BCI design is improved by using spectral power and asymmetry ratios from gamma (24-37 Hz) <span class="hlt">band</span> in <span class="hlt">addition</span> to the lower frequency <span class="hlt">bands</span>. In the experimental study, EEG signals extracted during five mental tasks from four subjects were used. Elman neural network (ENN) trained by the resilient backpropagation algorithm was used to classify the power and asymmetry ratios from EEG into different combinations of two mental tasks. The results indicated that ((1) the classification performance and training time of the BCI design were improved through the use of <span class="hlt">additional</span> gamma <span class="hlt">band</span> features; (2) classification performances were nearly invariant to the number of ENN hidden units or feature extraction method.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017PhRvE..96f2701H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017PhRvE..96f2701H"><span>Optical <span class="hlt">band</span> gap in a cholesteric elastomer doped by metallic nanospheres</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hernández, Julio C.; Reyes, J. Adrián</p> <p>2017-12-01</p> <p>We analyzed the optical <span class="hlt">band</span> gaps for axially propagating electromagnetic waves throughout a metallic doped cholesteric elastomer. The composed medium is made of metallic nanospheres (silver) randomly dispersed in a cholesteric elastomer liquid crystal whose dielectric properties can be represented by a resonant effective uniaxial tensor. We found that the <span class="hlt">band</span> gap properties of the periodic system greatly depend on the volume fraction of nanoparticles in the cholesteric elastomer. In particular, we observed a displacement of the reflection <span class="hlt">band</span> for quite small fraction volumes whereas for larger values of this fraction there appears a secondary <span class="hlt">band</span> in the higher frequency region. We also have calculated the transmittance and reflectance spectra for our system. These calculations verify the mentioned <span class="hlt">band</span> structure and provide <span class="hlt">additional</span> information about the polarization features of the radiation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015JAP...117c5704S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015JAP...117c5704S"><span>Evidence of ion intercalation mediated <span class="hlt">band</span> structure modification and opto-ionic coupling in lithium niobite</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Shank, Joshua C.; Tellekamp, M. Brooks; Doolittle, W. Alan</p> <p>2015-01-01</p> <p>The theoretically suggested <span class="hlt">band</span> structure of the novel p-type semiconductor lithium niobite (LiNbO2), the direct coupling of photons to ion motion, and optically induced <span class="hlt">band</span> structure modifications are investigated by temperature dependent photoluminescence. LiNbO2 has previously been used as a memristor material but is shown here to be useful as a sensor owing to the electrical, optical, and chemical ease of lithium removal and insertion. Despite the high concentration of vacancies present in lithium niobite due to the intentional removal of lithium atoms, strong photoluminescence spectra are observed even at room temperature that experimentally confirm the suggested <span class="hlt">band</span> structure implying transitions from a flat conduction <span class="hlt">band</span> to a degenerate valence <span class="hlt">band</span>. Removal of small amounts of lithium significantly modifies the photoluminescence spectra including <span class="hlt">additional</span> larger than stoichiometric-<span class="hlt">band</span> gap features. Sufficient removal of lithium results in the elimination of the photoluminescence response supporting the predicted transition from a direct to indirect <span class="hlt">band</span> gap semiconductor. In <span class="hlt">addition</span>, non-thermal coupling between the incident laser and lithium ions is observed and results in modulation of the electrical impedance.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20000073297&hterms=forest+trees&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3Dforest%2Btrees','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20000073297&hterms=forest+trees&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3Dforest%2Btrees"><span>Passive Microwave Measurements Over Conifer Forests at L-<span class="hlt">Band</span> and C-<span class="hlt">Band</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>LeVine, D. M.; Lang, R.; Chauhan, N.; Kim, E.; Bidwell, S.; Goodberlet, M.; Haken, M.; deMatthaeis, P.</p> <p>2000-01-01</p> <p>Measurements have been made at L-<span class="hlt">band</span> and C-<span class="hlt">band</span> over conifer forests in Virginia to study the response of passive microwave instruments to biomass and soil moisture. A series of aircraft measurements were made in July, August and November, 1999 over relatively homogenous conifer forests of varying biomass. Three radiometers participated in these measurements. These were: 1) the L-<span class="hlt">band</span> radiometer ESTAR, a horizontally polarized synthetic aperture radiometer which has been used extensively in past measurements of soil moisture; 2) the L-<span class="hlt">band</span> radiometer SLFMR, a vertically polarized cross-track scanner which has been used successfully in the past for mapping sea surface salinity; and 3) The ACMR, a new C-<span class="hlt">band</span> radiometer which operates at V- and H-polarization and in the configuration for these experiments did not scan. All three radiometers were flown on the NASA P-3 aircraft based at the Goddard Space Flight Center's Wallops Flight Facility. The ESTAR and SLFMR were mounted in the bomb bay of the P-3 and imaged across track whereas the ACMR was mounted to look aft at 54 degrees up from nadir. Data was collected at altitudes of 915 meters and 457 meters. The forests consisted of relatively homogeneous "managed" stands of conifer located near Waverly, Virginia. This is a relatively flat area about 30 miles southeast of Richmond, VA with numerous stands of trees being grown for the forestry industry. The stands selected for study consisted of areas of regrowth and mature stands of pine. In <span class="hlt">addition</span>, a small stand of very large trees was observed. Soil moisture sampling was done in each stand during the aircraft over flights. Data was collected on July 7, August 27, November 15 and November 30, 1999. Measurements were made with ESTAR on all days. The ACMR flew on the summer missions and the SLFMR was present only on the August 27 flight. Soil moisture varied from quite dry on July 7 to quite moist on November 30 (which was shortly after a period of rain). The microwave</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1991PhRvB..4312662A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1991PhRvB..4312662A"><span>Polaronlike vibrational <span class="hlt">bands</span> of molecular crystals with one-dimensional hydrogen-bond chains: N-methylacetamide</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Araki, Gako; Suzuki, Kazuaki; Nakayama, Hideyuki; Ishii, Kikujiro</p> <p>1991-05-01</p> <p>N-methylacetamide (NMA) crystal forms one-dimensional hydrogen-bond chains, which are similar to those in an acetanilide (ACN) crystal for which an unconventional vibrational <span class="hlt">band</span> accompanying the amide-I <span class="hlt">band</span> has been observed. Infrared spectra of NMA crystals show an <span class="hlt">additional</span> <span class="hlt">band</span> on the small-wave-number side of the amide-II <span class="hlt">band</span> as the temperature is lowered. There is a close resemblance between this <span class="hlt">band</span> and the <span class="hlt">band</span> of ACN. It is likely that these <span class="hlt">bands</span> appear by the same mechanism. The polaron model, which has been employed to explain the <span class="hlt">band</span> of ACN, was found to be applicable also to the case of NMA, although the main vibrational mode is amide I in ACN and amide II in NMA.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li class="active"><span>24</span></li> <li><a href="#" onclick='return showDiv("page_25");'>25</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_24 --> <div id="page_25" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li><a href="#" onclick='return showDiv("page_24");'>24</a></li> <li class="active"><span>25</span></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="481"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28828134','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28828134"><span>Prenatal and postnatal evaluation of polymicrogyria with <span class="hlt">band</span> heterotopia.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Nagaraj, Usha D; Hopkin, Robert; Schapiro, Mark; Kline-Fath, Beth</p> <p>2017-09-01</p> <p>The coexistence of <span class="hlt">band</span> heterotopia and polymicrogyria is extremely rare though it has been reported in the presence of corpus callosum anomalies and megalencephaly. We present prenatal and postnatal MRI findings of a rare case of diffuse cortical malformation characterized by polymicrogyria and <span class="hlt">band</span> heterotopia. Agenesis of the corpus callosum and megalencephaly were also noted. In <span class="hlt">addition</span>, bilateral closed-lip schizencephaly was identified on postnatal MRI, which has not been previously reported with this combination of imaging findings. Polymicrogyria with <span class="hlt">band</span> heterotopia can occur and can be diagnosed with fetal MRI. The coexistence of corpus callosum anomalies and megalencephaly comprises a rare phenotype that has been previously described, suggesting an underlying genetic abnormality.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4102904','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4102904"><span><span class="hlt">Energy</span> deposition by heavy ions: <span class="hlt">Additivity</span> of kinetic and potential <span class="hlt">energy</span> contributions in hillock formation on CaF2</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Wang, Y. Y.; Grygiel, C.; Dufour, C.; Sun, J. R.; Wang, Z. G.; Zhao, Y. T.; Xiao, G. Q.; Cheng, R.; Zhou, X. M.; Ren, J. R.; Liu, S. D.; Lei, Y.; Sun, Y. B.; Ritter, R.; Gruber, E.; Cassimi, A.; Monnet, I.; Bouffard, S.; Aumayr, F.; Toulemonde, M.</p> <p>2014-01-01</p> <p>Modification of surface and bulk properties of solids by irradiation with ion beams is a widely used technique with many applications in material science. In this study, we show that nano-hillocks on CaF2 crystal surfaces can be formed by individual impact of medium <span class="hlt">energy</span> (3 and 5 MeV) highly charged ions (Xe22+ to Xe30+) as well as swift (kinetic <span class="hlt">energies</span> between 12 and 58 MeV) heavy xenon ions. For very slow highly charged ions the appearance of hillocks is known to be linked to a threshold in potential <span class="hlt">energy</span> (Ep) while for swift heavy ions a minimum electronic <span class="hlt">energy</span> loss per unit length (Se) is necessary. With our results we bridge the gap between these two extreme cases and demonstrate, that with increasing <span class="hlt">energy</span> deposition via Se the Ep-threshold for hillock production can be lowered substantially. Surprisingly, both mechanisms of <span class="hlt">energy</span> deposition in the target surface seem to contribute in an <span class="hlt">additive</span> way, which can be visualized in a phase diagram. We show that the inelastic thermal spike model, originally developed to describe such material modifications for swift heavy ions, can be extended to the case where both kinetic and potential <span class="hlt">energies</span> are deposited into the surface. PMID:25034006</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25034006','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25034006"><span><span class="hlt">Energy</span> deposition by heavy ions: <span class="hlt">additivity</span> of kinetic and potential <span class="hlt">energy</span> contributions in hillock formation on CaF2.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wang, Y Y; Grygiel, C; Dufour, C; Sun, J R; Wang, Z G; Zhao, Y T; Xiao, G Q; Cheng, R; Zhou, X M; Ren, J R; Liu, S D; Lei, Y; Sun, Y B; Ritter, R; Gruber, E; Cassimi, A; Monnet, I; Bouffard, S; Aumayr, F; Toulemonde, M</p> <p>2014-07-18</p> <p>Modification of surface and bulk properties of solids by irradiation with ion beams is a widely used technique with many applications in material science. In this study, we show that nano-hillocks on CaF2 crystal surfaces can be formed by individual impact of medium <span class="hlt">energy</span> (3 and 5 MeV) highly charged ions (Xe(22+) to Xe(30+)) as well as swift (kinetic <span class="hlt">energies</span> between 12 and 58 MeV) heavy xenon ions. For very slow highly charged ions the appearance of hillocks is known to be linked to a threshold in potential <span class="hlt">energy</span> (Ep) while for swift heavy ions a minimum electronic <span class="hlt">energy</span> loss per unit length (Se) is necessary. With our results we bridge the gap between these two extreme cases and demonstrate, that with increasing <span class="hlt">energy</span> deposition via Se the Ep-threshold for hillock production can be lowered substantially. Surprisingly, both mechanisms of <span class="hlt">energy</span> deposition in the target surface seem to contribute in an <span class="hlt">additive</span> way, which can be visualized in a phase diagram. We show that the inelastic thermal spike model, originally developed to describe such material modifications for swift heavy ions, can be extended to the case where both kinetic and potential <span class="hlt">energies</span> are deposited into the surface.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010NTA.....1...69T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010NTA.....1...69T"><span>Fatigue level estimation of monetary bills based on frequency <span class="hlt">band</span> acoustic signals with feature selection by supervised SOM</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Teranishi, Masaru; Omatu, Sigeru; Kosaka, Toshihisa</p> <p></p> <p>Fatigued monetary bills adversely affect the daily operation of automated teller machines (ATMs). In order to make the classification of fatigued bills more efficient, the development of an automatic fatigued monetary bill classification method is desirable. We propose a new method by which to estimate the fatigue level of monetary bills from the feature-selected frequency <span class="hlt">band</span> acoustic <span class="hlt">energy</span> pattern of banking machines. By using a supervised self-organizing map (SOM), we effectively estimate the fatigue level using only the feature-selected frequency <span class="hlt">band</span> acoustic <span class="hlt">energy</span> pattern. Furthermore, the feature-selected frequency <span class="hlt">band</span> acoustic <span class="hlt">energy</span> pattern improves the estimation accuracy of the fatigue level of monetary bills by adding frequency domain information to the acoustic <span class="hlt">energy</span> pattern. The experimental results with real monetary bill samples reveal the effectiveness of the proposed method.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018BSRSL..87..321A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018BSRSL..87..321A"><span>Multi-<span class="hlt">band</span> optical variability studies of Blazars</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Agarwal, Aditi</p> <p>2018-04-01</p> <p>To search for optical variability on a wide range of timescales, we have carried out photometric monitoring of a dozen blazars. CCD magnitudes in B, V, R and I pass-<span class="hlt">bands</span> were determined for > 10,000f new optical observations from 300 nights made during 2011 – 2016, with an average length of 4 h each, using seven optical telescopes: four in Bulgaria, one in Greece, and two in India. We measured multiband optical flux and colour variations on diverse timescales. Blazar variability studies helped us in understanding their nature and extreme conditions within the emission region. To explain possible physical causes of the observed spectral variability, we also investigated spectral <span class="hlt">energy</span> distributions using B, V, R, I, J and K pass-<span class="hlt">band</span> data.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22304165-two-band-superlinear-electroluminescence-gasb-based-nanoheterostructures-alsb-inas-sub-sb-sub-alsb-deep-quantum-well','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22304165-two-band-superlinear-electroluminescence-gasb-based-nanoheterostructures-alsb-inas-sub-sb-sub-alsb-deep-quantum-well"><span>Two-<span class="hlt">band</span> superlinear electroluminescence in GaSb based nanoheterostructures with AlSb/InAs{sub 1−x} Sb{sub x}/AlSb deep quantum well</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Mikhailova, M. P.; Ivanov, E. V.; Danilov, L. V.</p> <p>2014-06-14</p> <p>We report on superlinear electroluminescent structures based on AlSb/InAs{sub 1−x}Sb{sub x}/AlSb deep quantum wells grown by MOVPE on n-GaSb:Te substrates. Dependence of the electroluminescence (EL) spectra and optical power on the drive current in nanoheterostructures with AlSb/InAs{sub 1−x}Sb{sub x}/AlSb quantum well at 77–300 K temperature range was studied. Intensive two-<span class="hlt">band</span> superlinear EL in the 0.5–0.8 eV photon <span class="hlt">energy</span> range was observed. Optical power enhancement with the increasing drive current at room temperature is caused by the contribution of the <span class="hlt">additional</span> electron-hole pairs due to the impact ionization by the electrons heated at the high <span class="hlt">energy</span> difference between AlSb and the first electronmore » level E{sub e1} in the InAsSb QW. Study of the EL temperature dependence at 90–300 K range enabled us to define the role of the first and second heavy hole levels in the radiative recombination process. It was shown that with the temperature decrease, the relation between the <span class="hlt">energies</span> of the valence <span class="hlt">band</span> offset and the second heavy hole <span class="hlt">energy</span> level changes due to the temperature transformation of the <span class="hlt">energy</span> <span class="hlt">band</span> diagram. That is the reason why the EL spectrum revealed radiative transitions from the first electron level E{sub e1} to the first hole level E{sub h1} in the whole temperature range (90–300 K), while the emission <span class="hlt">band</span> related with the transitions to the second hole level occurred only at T > 200 K. Comparative examination of the nanostructures with high <span class="hlt">band</span> offsets and different interface types (AlAs-like and InSb-like) reveals more intense EL and optical power enhancement at room temperature in the case of AlAs-like interface that could be explained by the better quality of the heterointerface and more efficient hole localization.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013PhDT.......432M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013PhDT.......432M"><span>Correlation of Photocatalytic Activity with <span class="hlt">Band</span> Structure of Low-dimensional Semiconductor Nanostructures</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Meng, Fanke</p> <p></p> <p>Photocatalytic hydrogen generation by water splitting is a promising technique to produce clean and renewable solar fuel. The development of effective semiconductor photocatalysts to obtain efficient photocatalytic activity is the key objective. However, two critical reasons prevent wide applications of semiconductor photocatalysts: low light usage efficiency and high rates of charge recombination. In this dissertation, several low-dimensional semiconductors were synthesized with hydrothermal, hydrolysis, and chemical impregnation methods. The <span class="hlt">band</span> structures of the low-dimensional semiconductor materials were engineered to overcome the above mentioned two shortcomings. In <span class="hlt">addition</span>, the correlation between the photocatalytic activity of the low-dimensional semiconductor materials and their <span class="hlt">band</span> structures were studied. First, we studied the effect of oxygen vacancies on the photocatalytic activity of one-dimensional anatase TiO2 nanobelts. Given that the oxygen vacancy plays a significant role in <span class="hlt">band</span> structure and photocatalytic performance of semiconductors, oxygen vacancies were introduced into the anatase TiO2 nanobelts during reduction in H2 at high temperature. The oxygen vacancies of the TiO2 nanobelts boosted visible-light-responsive photocatalytic activity but weakened ultraviolet-light-responsive photocatalytic activity. As oxygen vacancies are commonly introduced by dopants, these results give insight into why doping is not always beneficial to the overall photocatalytic performance despite increases in absorption. Second, we improved the photocatalytic performance of two-dimensional lanthanum titanate (La2Ti2 O7) nanosheets, which are widely studied as an efficient photocatalyst due to the unique layered crystal structure. Nitrogen was doped into the La2Ti2O7 nanosheets and then Pt nanoparticles were loaded onto the La2Ti2O7 nanosheets. Doping nitrogen narrowed the <span class="hlt">band</span> gap of the La2Ti 2O7 nanosheets by introducing a continuum of states by the valence</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017APS..MARH52004Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017APS..MARH52004Y"><span>Realization of space-time inversion-invariant topological semimetal-<span class="hlt">bands</span> in superconducting quantum circuits.</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yu, Y.; Tan, X.; Liu, Q.; Xue, G.; Yu, H.; Zhao, Y.; Wang, Z.</p> <p></p> <p>Topological <span class="hlt">band</span> theory has attracted much attention since several types of topological metals and semimetals have been explored. These robustness of nodal <span class="hlt">band</span> structures are symmetry-protected, whose topological features have deepened and widened the understandings of condensed matter physics. Meanwhile, as artificial quantum systems superconducting circuits possess high controllability, supplying a powerful approach to investigate topological properties of condensed matter systems. We realize a Hamiltonian with space-time (PT) symmetry by mapping momentum space of nodal <span class="hlt">band</span> structure to parameter space in a superconducting quantum circuit. By measuring <span class="hlt">energy</span> spectrum of the system, we observe the gapless <span class="hlt">band</span> structure of topological semimetals, shown as Dirac points in momentum space. The phase transition from topological semimetal to topological insulator can be realized by continuously tuning the parameter in Hamiltonian. We add perturbation to broken time reversal symmetry. As long as the combined PT symmetry is preserved, the Dirac points of the topological semimetal are still observable, suggesting the robustness of the topological protection of the gapless <span class="hlt">energy</span> <span class="hlt">band</span>. Our work open a platform to simulate the relation between the symmetry and topological stability in condensed matter systems. Supported by the NKRDP of China (2016YFA0301802) and the GRF of Hong Kong (HKU173051/14P&HKU173055/15P).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/21389492','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21389492"><span>The quasiparticle <span class="hlt">band</span> structure of zincblende and rocksalt ZnO.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Dixit, H; Saniz, R; Lamoen, D; Partoens, B</p> <p>2010-03-31</p> <p>We present the quasiparticle <span class="hlt">band</span> structure of ZnO in its zincblende (ZB) and rocksalt (RS) phases at the Γ point, calculated within the GW approximation. The effect of the p-d hybridization on the quasiparticle corrections to the <span class="hlt">band</span> gap is discussed. We compare three systems, ZB-ZnO which shows strong p-d hybridization and has a direct <span class="hlt">band</span> gap, RS-ZnO which is also hybridized but includes inversion symmetry and therefore has an indirect <span class="hlt">band</span> gap, and ZB-ZnS which shows a weaker hybridization due to a change of the chemical species from oxygen to sulfur. The quasiparticle corrections are calculated with different numbers of valence electrons in the Zn pseudopotential. We find that the Zn(20+) pseudopotential is essential for the adequate treatment of the exchange interaction in the self-<span class="hlt">energy</span>. The calculated GW <span class="hlt">band</span> gaps are 2.47 eV and 4.27 eV respectively, for the ZB and RS phases. The ZB-ZnO <span class="hlt">band</span> gap is underestimated compared to the experimental value of 3.27 by ∼ 0.8 eV. The RS-ZnO <span class="hlt">band</span> gap compares well with the experimental value of 4.5 eV. The underestimation for ZB-ZnO is correlated with the strong p-d hybridization. The GW <span class="hlt">band</span> gap for ZnS is 3.57 eV, compared to the experimental value of 3.8 eV.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20020017758','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20020017758"><span><span class="hlt">Band</span> Anticrossing in Highly Mismatched Compound Semiconductor Alloys</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Yu, Kin Man; Wu, J.; Walukiewicz, W.; Ager, J. W.; Haller, E. E.; Miotkowski, I.; Ramdas, A.; Su, Ching-Hua; Whitaker, Ann F. (Technical Monitor)</p> <p>2001-01-01</p> <p>Compound semiconductor alloys in which metallic anions are partially replaced with more electronegative isoelectronic atoms have recently attracted significant attention. Group IIIN(x)V(1-x), alloys with a small amount of the electronegative N substituting more metallic column V elements has been the most extensively studied class of such Highly Mismatched Alloys (HMAs). We have shown that many of the unusual properties of the IIIN(x),V(1-x) alloys can be well explained by the <span class="hlt">Band</span> Anticrossing (BAC) model that describes the electronic structure in terms of an interaction between highly localized levels of substitutional N and the extended states of the host semiconductor matrix. Most recently the BAC model has been also used to explain similar modifications of the electronic <span class="hlt">band</span> structure observed in Te-rich ZnS(x)Te(l-x) and ZnSe(Y)Te(1-y) alloys. To date studies of HMAs have been limited to materials with relatively small concentrations of highly electronegative atoms. Here we report investigations of the electronic structure of ZnSe(y)Te(1-y) alloys in the entire composition range, 0 less than or equal to y less than or equal to 1. The samples used in this study are bulk ZnSe(y)Te(1-y) crystals grown by either a modified Bridgman method or by physical vapor transport. Photomodulated reflection (PR) spectroscopy was used to measure the composition dependence of optical transitions from the valence <span class="hlt">band</span> edge and from the spin-orbit split off <span class="hlt">band</span> to the conduction <span class="hlt">band</span>. The pressure dependence of the <span class="hlt">band</span> gap was measured using optical absorption in a diamond anvil cell. We find that the <span class="hlt">energy</span> of the spin-orbit split off valence <span class="hlt">band</span> edge does not depend on composition and is located at about 3 eV below the conduction <span class="hlt">band</span> edge of ZnSe. On the Te-rich side the pressure and the composition dependence of the optical transitions are well explained by the BAC model which describes the downward shift of the conduction <span class="hlt">band</span> edge in terms of the interaction between</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2010-title47-vol1/pdf/CFR-2010-title47-vol1-sec15-223.pdf','CFR'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2010-title47-vol1/pdf/CFR-2010-title47-vol1-sec15-223.pdf"><span>47 CFR 15.223 - Operation in the <span class="hlt">band</span> 1.705-10 MHz.</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collectionCfr.action?selectedYearFrom=2010&page.go=Go">Code of Federal Regulations, 2010 CFR</a></p> <p></p> <p>2010-10-01</p> <p>... 47 Telecommunication 1 2010-10-01 2010-10-01 false Operation in the <span class="hlt">band</span> 1.705-10 MHz. 15.223 Section 15.223 Telecommunication FEDERAL COMMUNICATIONS COMMISSION GENERAL RADIO FREQUENCY DEVICES Intentional Radiators Radiated Emission Limits, <span class="hlt">Additional</span> Provisions § 15.223 Operation in the <span class="hlt">band</span> 1.705-10...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFM.B43B0582R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFM.B43B0582R"><span>A Passive Microwave L-<span class="hlt">Band</span> Boreal Forest Freeze/Thaw and Vegetation Phenology Study</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Roy, A.; Sonnentag, O.; Pappas, C.; Mavrovic, A.; Royer, A.; Berg, A. A.; Rowlandson, T. L.; Lemay, J.; Helgason, W.; Barr, A.; Black, T. A.; Derksen, C.; Toose, P.</p> <p>2016-12-01</p> <p>The boreal forest is the second largest land biome in the world and thus plays a major role in the global and regional climate systems. The extent, timing and duration of seasonal freeze/thaw (F/T) state influences vegetation developmental stages (phenology) and, consequently, constitute an important control on how boreal forest ecosystems exchange carbon, water and <span class="hlt">energy</span> with the atmosphere. The effective retrieval of seasonal F/T state from L-<span class="hlt">Band</span> radiometry was demonstrated using satellite mission. However, disentangling the seasonally differing contributions from forest overstory and understory vegetation, and the soil surface to the satellite signal remains challenging. Here we present initial results from a radiometer field campaign to improve our understanding of the L-<span class="hlt">Band</span> derived boreal forest F/T signal and vegetation phenology. Two L-<span class="hlt">Band</span> surface-based radiometers (SBR) are installed on a micrometeorological tower at the Southern Old Black Spruce site in central Saskatchewan over the 2016-2017 F/T season. One radiometer unit is installed on the flux tower so it views forest including all overstory and understory vegetation and the moss-covered ground surface. A second radiometer unit is installed within the boreal forest overstory, viewing the understory and the ground surface. The objectives of our study are (i) to disentangle the L-<span class="hlt">Band</span> F/T signal contribution of boreal forest overstory from the understory and ground surface, (ii) to link the L-<span class="hlt">Band</span> F/T signal to related boreal forest structural and functional characteristics, and (iii) to investigate the use of the L-<span class="hlt">Band</span> signal to characterize boreal forest carbon, water and <span class="hlt">energy</span> fluxes. The SBR observations above and within the forest canopy are used to retrieve the transmissivity (γ) and the scattering albedo (ω), two parameters that describe the emission of the forest canopy though the F/T season. These two forest parameters are compared with boreal forest structural and functional</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016PhRvB..94o5202G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016PhRvB..94o5202G"><span>Diluted magnetic semiconductors with narrow <span class="hlt">band</span> gaps</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gu, Bo; Maekawa, Sadamichi</p> <p>2016-10-01</p> <p>We propose a method to realize diluted magnetic semiconductors (DMSs) with p - and n -type carriers by choosing host semiconductors with a narrow <span class="hlt">band</span> gap. By employing a combination of the density function theory and quantum Monte Carlo simulation, we demonstrate such semiconductors using Mn-doped BaZn2As2 , which has a <span class="hlt">band</span> gap of 0.2 eV. In <span class="hlt">addition</span>, we found a nontoxic DMS Mn-doped BaZn2Sb2 , of which the Curie temperature Tc is predicted to be higher than that of Mn-doped BaZn2As2 , the Tc of which was up to 230 K in a recent experiment.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018MS%26E..327c2015C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018MS%26E..327c2015C"><span>Composite Gypsum Binders with Silica-containing <span class="hlt">Additives</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Chernysheva, N. V.; Lesovik, V. S.; Drebezgova, M. Yu; Shatalova, S. V.; Alaskhanov, A. H.</p> <p>2018-03-01</p> <p>New types of fine mineral <span class="hlt">additives</span> are proposed for designing water-resistant Composite Gypsum Binders (CGB); these <span class="hlt">additives</span> significantly differ from traditional quartz feed: wastes from wet magnetic separation of <span class="hlt">Banded</span> Iron Formation (BIF WMS waste), nanodispersed silica powder (NSP), chalk. Possibility of their combined use has been studied as well.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26911659','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26911659"><span>A self-sacrifice template route to iodine modified BiOIO3: <span class="hlt">band</span> gap engineering and highly boosted visible-light active photoreactivity.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Feng, Jingwen; Huang, Hongwei; Yu, Shixin; Dong, Fan; Zhang, Yihe</p> <p>2016-03-21</p> <p>The development of high-performance visible-light photocatalysts with a tunable <span class="hlt">band</span> gap has great significance for enabling wide-<span class="hlt">band</span>-gap (WBG) semiconductors visible-light sensitive activity and precisely tailoring their optical properties and photocatalytic performance. In this work we demonstrate the continuously adjustable <span class="hlt">band</span> gap and visible-light photocatalysis activation of WBG BiOIO3via iodine surface modification. The iodine modified BiOIO3 was developed through a facile in situ reduction route by applying BiOIO3 as the self-sacrifice template and glucose as the reducing agent. By manipulating the glucose concentration, the <span class="hlt">band</span> gap of the as-prepared modified BiOIO3 could be orderly narrowed by generation of the impurity or defect <span class="hlt">energy</span> level close to the conduction <span class="hlt">band</span>, thus endowing it with a visible light activity. The photocatalytic assessments uncovered that, in contrast to pristine BiOIO3, the modified BiOIO3 presents significantly boosted photocatalytic properties for the degradation of both liquid and gaseous contaminants, including Rhodamine B (RhB), methyl orange (MO), and ppb-level NO under visible light. <span class="hlt">Additionally</span>, the <span class="hlt">band</span> structure evolution as well as photocatalysis mechanism triggered by the iodine surface modification is investigated in detail. This study not only provides a novel iodine surface-modified BiOIO3 for environmental application, but also provides a facile and general way to develop highly efficient visible-light photocatalysts.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016IJMPB..3042004S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016IJMPB..3042004S"><span>Strong coupling diagram technique for the three-<span class="hlt">band</span> Hubbard model</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sherman, A.</p> <p>2016-03-01</p> <p>Strong coupling diagram technique equations are derived for hole Green’s functions of the three-<span class="hlt">band</span> Hubbard model, which describes Cu-O planes of high-Tc cuprates. The equations are self-consistently solved in the approximation, in which the series for the irreducible part in powers of the oxygen-copper hopping constant is truncated to two lowest-order terms. For parameters used for hole-doped cuprates, the calculated <span class="hlt">energy</span> spectrum consists of lower and upper Hubbard subbands of predominantly copper nature, oxygen <span class="hlt">bands</span> with a small admixture of copper states and the Zhang-Rice states of mixed nature, which are located between the lower Hubbard subband and oxygen <span class="hlt">bands</span>. The spectrum contains also pseudogaps near transition frequencies of Hubbard atoms on copper sites.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20150001286','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20150001286"><span>SNPP VIIRS Spectral <span class="hlt">Bands</span> Co-Registration and Spatial Response Characterization</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Lin, Guoqing; Tilton, James C.; Wolfe, Robert E.; Tewari, Krishna P.; Nishihama, Masahiro</p> <p>2013-01-01</p> <p>The Visible Infrared Imager Radiometer Suite (VIIRS) instrument onboard the Suomi National Polar-orbiting Partnership (SNPP) satellite was launched on 28 October 2011. The VIIRS has 5 imagery spectral <span class="hlt">bands</span> (I-<span class="hlt">bands</span>), 16 moderate resolution spectral <span class="hlt">bands</span> (M-<span class="hlt">bands</span>) and a panchromatic day/night <span class="hlt">band</span> (DNB). Performance of the VIIRS spatial response and <span class="hlt">band-to-band</span> co-registration (BBR) was measured through intensive pre-launch tests. These measurements were made in the non-aggregated zones near the start (or end) of scan for the I-<span class="hlt">bands</span> and M-<span class="hlt">bands</span> and for a limited number of aggregation modes for the DNB in order to test requirement compliance. This paper presents results based on a recently re-processed pre-launch test data. Sensor (detector) spatial impulse responses in the scan direction are parameterized in terms of ground dynamic field of view (GDFOV), horizontal spatial resolution (HSR), modulation transfer function (MTF), ensquared <span class="hlt">energy</span> (EE) and integrated out-of-pixel (IOOP) spatial response. Results are presented for the non-aggregation, 2-sample and 3-sample aggregation zones for the I-<span class="hlt">bands</span> and M-<span class="hlt">bands</span>, and for a limited number of aggregation modes for the DNB. On-orbit GDFOVs measured for the 5 I-<span class="hlt">bands</span> in the scan direction using a straight bridge are also presented. <span class="hlt">Band-to-band</span> co-registration (BBR) is quantified using the prelaunch measured <span class="hlt">band-to-band</span> offsets. These offsets may be expressed as fractions of horizontal sampling intervals (HSIs), detector spatial response parameters GDFOV or HSR. BBR bases on HSIs in the non-aggregation, 2-sample and 3-sample aggregation zones are presented. BBR matrices based on scan direction GDFOV and HSR are compared to the BBR matrix based on HSI in the non-aggregation zone. We demonstrate that BBR based on GDFOV is a better representation of footprint overlap and so this definition should be used in BBR requirement specifications. We propose that HSR not be used as the primary image quality indicator, since we</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017MRE.....4a5101B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017MRE.....4a5101B"><span>Experimental and theoretical investigation of relative optical <span class="hlt">band</span> gaps in graphene generations</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bhatnagar, Deepika; Singh, Sukhbir; Yadav, Sriniwas; Kumar, Ashok; Kaur, Inderpreet</p> <p>2017-01-01</p> <p>Size and chemical functionalization dependant optical <span class="hlt">band</span> gaps in graphene family nanomaterials were investigated by experimental and theoretical study using Tauc plot and density functional theory (DFT). We have synthesized graphene oxide through a modified Hummer’s method using graphene nanoplatelets and sequentially graphene quantum dots through hydrothermal reduction. The experimental results indicate that the optical <span class="hlt">band</span> gap in graphene generations was altered by reducing the size of graphene sheets and attachment of chemical functionalities like epoxy, hydroxyl and carboxyl groups plays a crucial role in varying optical <span class="hlt">band</span> gaps. It is further confirmed by DFT calculations that the π orbitals were more dominatingly participating in transitions shown by projected density of states and the molecular <span class="hlt">energy</span> spectrum represented the effect of attached functional groups along with discreteness in <span class="hlt">energy</span> levels. Theoretical results were found to be in good agreement with experimental results. All of the above different variants of graphene can be used in native or modified form for sensor design and optoelectronic applications.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19790038523&hterms=merkel&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dmerkel','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19790038523&hterms=merkel&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dmerkel"><span>NS001MS - Landsat-D thematic mapper <span class="hlt">band</span> aircraft scanner</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Richard, R. R.; Merkel, R. F.; Meeks, G. R.</p> <p>1978-01-01</p> <p>The thematic mapper is a multispectral scanner which will be launched aboard Landsat-D in the early 1980s. Compared with previous Landsat scanners, this instrument will have an improved spatial resolution (30 m) and new spectral <span class="hlt">bands</span>. Designated NS001MS, the scanner is designed to duplicate the thematic mapper spectral <span class="hlt">bands</span> plus two <span class="hlt">additional</span> <span class="hlt">bands</span> (1.0 to 1.3 microns and 2.08 to 2.35 microns) in an aircraft scanner for evaluation and investigation prior to design and launch of the final thematic mapper. Applicable specifications used in defining the thematic mapper were retained in the NS001MS design, primarily with respect to spectral bandwidths, noise equivalent reflectance, and noise equivalent difference temperature. The technical design and operational characteristics of the multispectral scanner (with thematic mapper <span class="hlt">bands</span>) are discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20110015306','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20110015306"><span>Radiometric Quality of the MODIS <span class="hlt">Bands</span> at 667 and 678nm</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Meister, Gerhard; Franz, Bryan A.</p> <p>2010-01-01</p> <p>The MODIS instruments on Terra and Aqua were designed to allow the measurement of chlorophyll fluorescence effects over ocean. The retrieval algorithm is based on the difference between the water-leaving radiances at 667nm and 678nm. The water-leaving radiances at these wavelengths are usually very low relative to the top- of-atmosphere radiances. The high radiometric accuracy needed to retrieve the small fluorescence signal lead to a dual gain design for the 667 and 678nm <span class="hlt">bands</span>. This paper discusses the benefits obtained from this design choice and provides justification for the use of only one set of gains for global processing of ocean color products. Noise characteristics of the two <span class="hlt">bands</span> and their related products are compared to other products of <span class="hlt">bands</span> from 412nm to 2130nm. The impact of polarization on the two <span class="hlt">bands</span> is discussed. In <span class="hlt">addition</span>, the impact of stray light on the two <span class="hlt">bands</span> is compared to other MODIS <span class="hlt">bands</span>.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li><a href="#" onclick='return showDiv("page_24");'>24</a></li> <li class="active"><span>25</span></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_25 --> <div class="footer-extlink text-muted" style="margin-bottom:1rem; text-align:center;">Some links on this page may take you to non-federal websites. Their policies may differ from this site.</div> </div><!-- container --> <footer><a id="backToTop" href="#top"> </a><nav><a id="backToTop" href="#top"> </a><ul class="links"><a id="backToTop" href="#top"> </a><li><a id="backToTop" href="#top"></a><a href="/sitemap.html">Site Map</a></li> <li><a href="/members/index.html">Members Only</a></li> <li><a href="/website-policies.html">Website Policies</a></li> <li><a href="https://doe.responsibledisclosure.com/hc/en-us" target="_blank">Vulnerability Disclosure Program</a></li> <li><a href="/contact.html">Contact Us</a></li> </ul> <div class="small">Science.gov is maintained by the U.S. Department of Energy's <a href="https://www.osti.gov/" target="_blank">Office of Scientific and Technical Information</a>, in partnership with <a href="https://www.cendi.gov/" target="_blank">CENDI</a>.</div> </nav> </footer> <script type="text/javascript"><!-- // var lastDiv = ""; function showDiv(divName) { // hide last div if (lastDiv) { document.getElementById(lastDiv).className = "hiddenDiv"; } //if value of the box is not nothing and an object with that name exists, then change the class if (divName && document.getElementById(divName)) { document.getElementById(divName).className = "visibleDiv"; lastDiv = divName; } } //--> </script> <script> /** * Function that tracks a click on an outbound link in Google Analytics. * This function takes a valid URL string as an argument, and uses that URL string * as the event label. */ var trackOutboundLink = function(url,collectionCode) { try { h = window.open(url); setTimeout(function() { ga('send', 'event', 'topic-page-click-through', collectionCode, url); }, 1000); } catch(err){} }; </script> <!-- Google Analytics --> <script> (function(i,s,o,g,r,a,m){i['GoogleAnalyticsObject']=r;i[r]=i[r]||function(){ (i[r].q=i[r].q||[]).push(arguments)},i[r].l=1*new Date();a=s.createElement(o), m=s.getElementsByTagName(o)[0];a.async=1;a.src=g;m.parentNode.insertBefore(a,m) })(window,document,'script','//www.google-analytics.com/analytics.js','ga'); ga('create', 'UA-1122789-34', 'auto'); ga('send', 'pageview'); </script> <!-- End Google Analytics --> <script> showDiv('page_1') </script> </body> </html>