Sample records for kev energy bands

  1. 950 keV X-Band Linac For Material Recognition Using Two-Fold Scintillator Detector As A Concept Of Dual-Energy X-Ray System

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    Lee, Kiwoo; Natsui, Takuya; Hirai, Shunsuke

    2011-06-01

    One of the advantages of applying X-band 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-band Linac and achieved maximum X-ray energy 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-band system. In addition, we have designed two-fold scintillator detector in dual energy 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-band 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

  2. Hard x-ray broad band Laue lenses (80-600 keV): building methods and performances

    NASA Astrophysics Data System (ADS)

    Virgilli, E.; Frontera, F.; Rosati, P.; Liccardo, V.; Squerzanti, S.; Carassiti, V.; Caroli, E.; Auricchio, N.; Stephen, J. B.

    2015-09-01

    We present the status of the LAUE project devoted to develop a technology for building a 20 meter long focal length Laue lens for hard X-/soft gamma-ray astronomy (80-600 keV). The Laue lens is composed of bent crystals of Gallium Arsenide (GaAs, 220) and Germanium (Ge, 111), and, for the first time, the focusing property of bent crystals has been exploited for this field of applications. We show the preliminary results concerning the adhesive employed to fix the crystal tiles over the lens support, the positioning accuracy obtained and possible further improvements. The Laue lens petal that will be completed in a few months has a pass band of 80-300 keV and is a fraction of an entire Laue lens capable of focusing x-rays up to 600 keV, possibly extendable down to ~20-30 keV with suitable low absorption crystal materials and focal length. The final goal is to develop a focusing optics that can improve the sensitivity over current telescopes in this energy band by 2 orders of magnitude.

  3. Simbol-X: a formation flight mission with an unprecedented imaging capability in the 0.5-80 keV energy band

    NASA Astrophysics Data System (ADS)

    Tagliaferri, Gianpiero; Ferrando, Philippe; Le Duigou, Jean-Michel; Pareschi, Giovanni; Laurent, Philippe; Malaguti, Giuseppe; Clédassou, Rodolphe; Piermaria, Mauro; La Marle, Olivier; Fiore, Fabrizio; Giommi, Paolo

    2017-11-01

    The discovery of X-ray emission from cosmic sources in the 1960s has opened a new powerful observing window on the Universe. In fact, the exploration of the X-ray sky during the 70s-90s has established X-ray astronomy as a fundamental field of astrophysics. Today, the emission from astrophysical sources is by large best known at energies below 10 keV. The main reason for this situation is purely technical since grazing incidence reflection has so far been limited to the soft X-ray band. Above 10 keV all the observations have been obtained with collimated detectors or coded mask instruments. To make a leap step forward in Xray astronomy above 10 keV it is necessary to extend the principle of focusing X ray optics to higher energies, up to 80 keV and beyond. To this end, ASI and CNES are presently studying the implementation of a X-ray mission called Simbol-X. Taking advantage of emerging technology in mirror manufacturing and spacecraft formation flying, Simbol-X will push grazing incidence imaging up to 80 keV and beyond, providing a strong improvement both in sensitivity and angular resolution compared to all instruments that have operated so far above 10 keV. This technological breakthrough will open a new highenergy window in astrophysics and cosmology. Here we will address the problematic of the development for such a distributed and deformable instrument. We will focus on the main performances of the telescope, like angular resolution, sensitivity and source localization. We will also describe the specificity of the calibration aspects of the payload distributed over two satellites and therefore in a not "frozen" configuration.

  4. Neutron fluence and energy reconstruction with the IRSN recoil detector μ-TPC at 27 keV, 144 keV and 565 keV

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    Maire, D.; Lebreton, L.; Richer, J.P.

    2015-07-01

    The French Institute for Radioprotection and Nuclear Safety (IRSN), associated to the French Metrology Institute (LNE), is developing a time projection chamber using a Micromegas anode: μ-TPC. This work is carried out in collaboration with the Laboratory of Subatomic Physics and Cosmology (LPSC). The aim is to characterize with a primary procedure the energy distribution of neutron fluence in the energy range 8 keV - 1 MeV. The time projection chambers are gaseous detectors, which are able to measure charged particles energy and to reconstruct their track if a pixelated anode is used. In our case, the gas is usedmore » as a (n, p) converter in order to detect neutrons down to few keV. Coming from elastic collisions with neutrons, recoil protons lose a part of their kinetic energy by ionizing the gas. The ionization electrons are drifted toward a pixelated anode (2D projection), read at 50 MHz by a self-triggered electronic system to obtain the third track dimension. The neutron energy is reconstructed event by event thanks to proton scattering angle and proton energy measurements. The scattering angle is deduced from the 3D track. The proton energy is obtained by charge collection measurements, knowing the ionization quenching factor (i.e. the part of proton kinetic energy lost by ionizing the gas). The fluence is calculated thanks to the detected events number and the simulated detector response. The μ-TPC is a new reliable detector which enables to measure energy distribution of the neutron fluence without deconvolution or neutron calibration contrary to usual gaseous counters. The μ-TPC is still being developed and measurements have been carried out at the AMANDE facility, with neutrons energies going from 8 keV to 565 keV. After the context and the μ-TPC working principle presentation, measurements of the neutron energy and fluence at 27.2 keV, 144 keV and 565 keV are shown and compared to the complete detector simulation. This work shows the first direct

  5. ANALYSIS OF ENERGY LOSSES OF A 30-kev ELECTRON BEAM IN THE FLUORIDE, CHLORIDE, AND BROMIDE OF LITHIUM (in French)

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    Pradal, F.; Gout, C.

    1963-02-01

    The energy loss of a 30-kev electron beam in films of LiF, LiCl, and LiBr were analyzed with a magnetic spectrograph. For LiF, the results are compared to the absorption curve in the ultraviolet. The rays observed seem due to the excitation of valence band electrons of the 2s band of F/sup -/ and the 1s band of Li/sup +/. In some cases, energy losses less than 10 ev were observed, which seems connected to the presence of color centers. (tr-auth)

  6. Sub-second variations of high energy ( 300 keV) hard X-ray emission from solar flares

    NASA Technical Reports Server (NTRS)

    Bai, Taeil

    1986-01-01

    Subsecond variations of hard X-ray emission from solar flares were first observed with a balloon-borne detector. With the launch of the Solar Maximum Mission (SMM), it is now well known that subsecond variations of hard X-ray emission occur quite frequently. Such rapid variations give constraints on the modeling of electron energization. Such rapid variations reported until now, however, were observed at relatively low energies. Fast mode data obtained by the Hard X-ray Burst Spectrometer (HXRBS) has time resolution of approximately 1 ms but has no energy resolution. Therefore, rapid fluctuations observed in the fast-mode HXRBS data are dominated by the low energy hard X-rays. It is of interest to know whether rapid fluctuations are observed in high-energy X-rays. The highest energy band at which subsecond variations were observed is 223 to 1057 keV. Subsecond variations observed with HXRBS at energies greater than 300 keV are reported, and the implications discussed.

  7. Energy dependence of the band-limited noise in black hole X-ray binaries★

    NASA Astrophysics Data System (ADS)

    Stiele, H.; Yu, W.

    2015-10-01

    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 band-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 band-limited noise component and study covariance spectra. For observations that show a noise component with a characteristic frequency above 1 Hz in the hard energy band (4-8 keV), we found this very same component at a lower frequency in the soft band (1-2 keV). This difference in characteristic frequency is an indication that while both the soft and the hard band photons contribute to the same band-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 energies, 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 energy dependence within the picture of energy-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 band is becoming significant.

  8. Hyper-filter-fluorescer spectrometer for x-rays above 120 keV

    DOEpatents

    Wang, Ching L.

    1983-01-01

    An apparatus utilizing filter-fluorescer combinations is provided to measure short bursts of high fluence x-rays above 120 keV energy, where there are no practical absorption edges available for conventional filter-fluorescer techniques. The absorption edge of the prefilter is chosen to be less than that of the fluorescer, i.e., E.sub.PRF E.sub.F. In this way, the response function is virtually zero between E.sub.PRF and E.sub.F and well defined and enhanced in an energy band of less than 1000 keV above the 120 keV energy.

  9. Solar neutrino flux at keV energies

    NASA Astrophysics Data System (ADS)

    Vitagliano, Edoardo; Redondo, Javier; Raffelt, Georg

    2017-12-01

    We calculate the solar neutrino and antineutrino flux in the keV energy range. The dominant thermal source processes are photo production (γ e→ e νbar nu), bremsstrahlung (e+Ze→ Ze+e+νbar nu), plasmon decay (γ→νbar nu), and νbar nu emission in free-bound and bound-bound transitions of partially ionized elements heavier than hydrogen and helium. These latter processes dominate in the energy range of a few keV and thus carry information about the solar metallicity. To calculate their rate we use libraries of monochromatic photon radiative opacities in analogy to a previous calculation of solar axion emission. Our overall flux spectrum and many details differ significantly from previous works. While this low-energy flux is not measurable with present-day technology, it could become a significant background for future direct searches for keV-mass sterile neutrino dark matter.

  10. The average 0.5-200 keV spectrum of local active galactic nuclei and a new determination of the 2-10 keV luminosity function at z ≈ 0

    NASA Astrophysics Data System (ADS)

    Ballantyne, D. R.

    2014-01-01

    The broad-band X-ray spectra of active galactic nuclei (AGNs) contains information about the nuclear environment from Schwarzschild radii scales (where the primary power law is generated in a corona) to distances of ˜1 pc (where the distant reflector may be located). In addition, the average shape of the X-ray spectrum is an important input into X-ray background synthesis models. Here, local (z ≈ 0) AGN luminosity functions (LFs) in five energy bands are used as a low-resolution, luminosity-dependent X-ray spectrometer in order to constrain the average AGN X-ray spectrum between 0.5 and 200 keV. The 15-55 keV LF measured by Swift-BAT is assumed to be the best determination of the local LF, and then a spectral model is varied to determine the best fit to the 0.5-2 keV, 2-10 keV, 3-20 keV and 14-195 keV LFs. The spectral model consists of a Gaussian distribution of power laws with a mean photon-index <Γ> and cutoff energy Ecut, as well as contributions from distant and disc reflection. The reflection strength is parametrized by varying the Fe abundance relative to solar, AFe, and requiring a specific Fe Kα equivalent width (EW). In this way, the presence of the X-ray Baldwin effect can be tested. The spectral model that best fits the four LFs has <Γ> = 1.85 ± 0.15, E_{cut}=270^{+170}_{-80} keV, A_{Fe}=0.3^{+0.3}_{-0.15}. The sub-solar AFe is unlikely to be a true measure of the gas-phase metallicity, but indicates the presence of strong reflection given the assumed Fe Kα EW. Indeed, parametrizing the reflection strength with the R parameter gives R=1.7^{+1.7}_{-0.85}. There is moderate evidence for no X-ray Baldwin effect. Accretion disc reflection is included in the best-fitting model, but it is relatively weak (broad iron Kα EW < 100 eV) and does not significantly affect any of the conclusions. A critical result of our procedure is that the shape of the local 2-10 keV LF measured by HEAO-1 and MAXI is incompatible with the LFs measured in the hard X

  11. Low-temperature volume radiation annealing of cold-worked bands of Al-Li-Cu-Mg alloy by 20-40 keV Ar+ ion

    NASA Astrophysics Data System (ADS)

    Ovchinnikov, V. V.; Gushchina, N. V.; Mozharovsky, S. M.; Kaigorodova, L. I.

    2017-01-01

    The processes of radiation-dynamic nature (in contrast to the thermally-activated processes) in the course of short-term irradiation of 1 mm thick bands of cold-worked aluminum alloy 1441 (of system Al-Li-Cu-Mg) with Ar+ 20-40 keV were studied. An effect of in-the-bulk (throughout the whole of metal bands thickness) low-temperature radiation annealing of the named alloy, multiply accelerated as compared with common thermal annealing processes was registered (with projected ranges of ions of considered energies definitely not exceeding 0.1 μm). The processes of recrystallization and intermetallic structure changes (occurring within a few seconds of Ar+ irradiation) have the common features as well as the differences in comparison with the results of two hour standard thermal annealing.

  12. The Hard X-ray 20-40 keV AGN Luminosity Function

    NASA Technical Reports Server (NTRS)

    Beckmann, V.; Soldi, S.; Shrader, C. R.; Gehrels, N.; Produit, N.

    2006-01-01

    We have compiled a complete, significance limited extragalactic sample based on approximately 25,000 deg(sup 2) to a limiting flux of 3 x 10(exp -11) ergs per square centimeter per second. (approximately 7,000 deg(sup 2)) to a flux limit of 10(exp -11) ergs per square centimeter per second)) in the 20 - 40 keV band with INTEGRAL. We have constructed a detailed exposure map to compensate for effects of non-uniform exposure. The flux-number relation is best described by a power-law with a slope of alpha = 1.66 plus or minus 0.11. The integration of the cumulative flux per unit area leads to f(sub 20-40 keV) = 2.6 x 10(exp -10) ergs per square centimeter per second per sr(sup -1) which is about 1% of the known 20-40 keV X-ray background. We present the first luminosity function of AGN in the 20-40 keV energy range, based on 68 extragalactic objects detected by the imager IBIS/ISGRI on-board INTEGRAL. The luminosity function shows a smoothly connected two power-law form, with an index of gamma (sub 1) = 0.9 below, and gamma (sub 2) = 2.2 above the turn-over luminosity of L(sub *), = 4.6 x 10(sup 43) ergs per second. The emissivity of all INTEGRAL AGNs per unit volume is W(sub 20-40keV)(greater than 10(sup 41) ergs per second) = 2.8 x 10(sup 38) ergs per second h(sup 3)(sub 70) Mpc(sup -3). These results are consistent with those derived in the 2-20keV energy band and do not show a significant contribution by Compton-thick objects. Because the sample used in this study is truly local (z(raised bar) = 0.022)), only limited conclusions can be drawn for the evolution of AGNs in this energy band. But the objects explaining the peak in the cosmic X-ray background are likely to be either low luminosity AGN (L(sub x) less than 10(sup 41) ergs per second) or of other type, such as intermediate mass black holes, clusters, and star forming regions.

  13. Structural analysis and evaluation of actual PC bridge using 950 keV/3.95 MeV X-band linacs

    NASA Astrophysics Data System (ADS)

    Takeuchi, H.; Yano, R.; Ozawa, I.; Mitsuya, Y.; Dobashi, K.; Uesaka, M.; Kusano, J.; Oshima, Y.; Ishida, M.

    2017-07-01

    In Japan, bridges constructed during the strong economic growth era are facing an aging problem and advanced maintenance methods have become strongly required recently. To meet this demand, we develop the on-site inspection system using 950 keV/3.95 MeV X-band (9.3 GHz) linac X-ray sources. These systems can visualize in seconds the inner states of bridges, including cracks of concrete, location and state of tendons (wires) and other imperfections. At the on-site inspections, 950 keV linac exhibited sufficient performance. But, for thicker concrete, it is difficult to visualize the internal state by 950 keV linac. Therefore, we proceeded the installation of 3.95 MeV linac for on-site bridge inspection. In addition, for accurate evaluation, verification on the parallel motion CT technique and FEM analysis are in progress.

  14. Experimental Determination of the HPGe Spectrometer Efficiency Calibration Curves for Various Sample Geometry for Gamma Energy from 50 keV to 2000 keV

    NASA Astrophysics Data System (ADS)

    Saat, Ahmad; Hamzah, Zaini; Yusop, Mohammad Fariz; Zainal, Muhd Amiruddin

    2010-07-01

    Detection efficiency of a gamma-ray spectrometry system is dependent upon among others, energy, sample and detector geometry, volume and density of the samples. In the present study the efficiency calibration curves of newly acquired (August 2008) HPGe gamma-ray spectrometry system was carried out for four sample container geometries, namely Marinelli beaker, disc, cylindrical beaker and vial, normally used for activity determination of gamma-ray from environmental samples. Calibration standards were prepared by using known amount of analytical grade uranium trioxide ore, homogenized in plain flour into the respective containers. The ore produces gamma-rays of energy ranging from 53 keV to 1001 keV. Analytical grade potassium chloride were prepared to determine detection efficiency of 1460 keV gamma-ray emitted by potassium isotope K-40. Plots of detection efficiency against gamma-ray energy for the four sample geometries were found to fit smoothly to a general form of ɛ = AΕa+BΕb, where ɛ is efficiency, Ε is energy in keV, A, B, a and b are constants that are dependent on the sample geometries. All calibration curves showed the presence of a "knee" at about 180 keV. Comparison between the four geometries showed that the efficiency of Marinelli beaker is higher than cylindrical beaker and vial, while cylindrical disk showed the lowest.

  15. Muonium formation at keV energies

    NASA Astrophysics Data System (ADS)

    Prokscha, T.; Morenzoni, E.; Garifianov, N.; Glückler, H.; Khasanov, R.; Luetkens, H.; Suter, A.

    2003-02-01

    Charge differentiation in μ + or muonium (Mu) as a consequence of the slowing down of μ + in matter is of fundamental interest in the μSR method. It is also of relevance for understanding the moderation process of μ + in van der Waals solids like s-Ne, s-Ar or s-N 2, which are the most suitable materials to generate epithermal μ + serving as a source for low-energy μ + (LE-μ +) beams. The LE-μ + beam at the Paul Scherrer Institut (PSI) allows us to study the formation of Mu at low implantation energies (0.5- 30 keV) in insulators and semiconductors. These investigations may help to clarify the interaction between the μ + and the electrons of its ionisation track since the number of track electrons can be tuned by varying the implantation energy. We present the first results of LE-μSR investigations on thin van der Waals solids (s-Ar, s-Xe), fused quartz ( SUPRASIL) and a quartz crystal, where the μ + and Mu fractions were measured as a function of the implantation energy.

  16. The Solar Flare 4: 10 keV X-ray Spectrum

    NASA Technical Reports Server (NTRS)

    Phillips, K. J. H.

    2004-01-01

    The 4-10 keV solar flare spectrum includes highly excited lines of stripped Ca, Fe, and Ni ions as well as a continuum steeply falling with energy. Groups of lines at approximately 7 keV and approximately 8 keV, observed during flares by the broad-band RHESSI spectrometer and called here the Fe-line and Fe/Ni-line features, are formed mostly of Fe lines but with Ni lines contributing to the approximately 8 keV feature. Possible temperature indicators of these line features are discussed - the peak or centroid energies of the Fe-line feature, the line ratio of the Fe-line to the Fe/Ni-line features, and the equivalent width of the Fe-line feature. The equivalent width is by far the most sensitive to temperature. However, results will be confused if, as is commonly believed, the abundance of Fe varies from flare to flare, even during the course of a single flare. With temperature determined from the thermal continuum, the Fe-line feature becomes a diagnostic of the Fe abundance in flare plasmas. These results are of interest for other hot plasmas in coronal ionization equilibrium such as stellar flare plasmas, hot gas in galaxies, and older supernova remnants.

  17. High-resolution integrated germanium Compton polarimeter for the γ-ray energy range 80 keV-1 MeV

    NASA Astrophysics Data System (ADS)

    Sareen, R. A.; Urban, W.; Barnett, A. R.; Varley, B. J.

    1995-06-01

    Parameters which govern the choice of a detection system to measure the linear polarization of γ rays at low energies are discussed. An integrated polarimeter is described which is constructed from a single crystal of germanium. It is a compact planar device with the sectors defined electrically, and which gives an energy resolution in the add-back mode of 1 keV at 300 keV. Its performance is demonstrated in a series of calibration measurements using both unpolarized radiation from radioactive sources and polarized γ rays from the 168Er(α,2n)170Yb reaction at Eα=25 MeV. Polarization measurements at energies as low as 84 keV have been achieved, where the sensitivity was 0.32±0.09. The sensitivity, efficiency, and energy resolution are reported. Our results indicate that energy resolution should be included in the definition of the figure of merit and we relate the new definition to earlier work. The comparisons show the advantages of the present design in the energy range below 300 keV and its competitiveness up to 1500 keV.

  18. Band structures in near spherical 138Ce

    NASA Astrophysics Data System (ADS)

    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.

    2009-06-01

    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 band 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 energy of 23 ℏ and 9511.3 keV, respectively, by including 53 new transitions. The negative parity ΔI=1 band, developed on the 6536.3 keV 15 level, has been conjectured to be a magnetic rotation band following a semiclassical analysis and comparing the systematics of similar bands in the neighboring nuclei. The said band is proposed to have a four quasiparticle configuration of [πgh]⊗[. Other band 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.

  19. Observations of solar flare photon energy spectra from 20 keV to 7 MeV

    NASA Technical Reports Server (NTRS)

    Yoshimori, M.; Watanabe, H.; Nitta, N.

    1985-01-01

    Solar flare photon energy spectra in the 20 keV to 7 MeV range are derived from the Apr. 1, Apr. 4, apr. 27 and May 13, 1981 flares. The flares were observed with a hard X-ray and a gamma-ray spectrometers on board the Hinotori satellite. The results show that the spectral shape varies from flare to flare and the spectra harden in energies above about 400 keV. Effects of nuclear line emission on the continuum and of higher energy electron bremsstrahlung are considered to explain the spectral hardening.

  20. Neutron Fluence and Energy Reconstruction with the LNE-IRSN/MIMAC Recoil Detector MicroTPC at 27 keV

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    Maire, D.; Lebreton, L.; Querre, Ph.

    2015-07-01

    The French Institute for Radiation protection and Nuclear Safety (IRSN), designated by the French Metrology Institute (LNE) for neutron metrology, is developing a time projection chamber using a Micromegas anode: microTPC. This work is carried out in collaboration with the Laboratory of Subatomic Physics and Cosmology (LPSC). The aim is to characterize the energy distribution of neutron fluence in the energy range 8 keV - 5 MeV with a primary procedure. The time projection chambers are gaseous detectors able to measure charged particles energy and to reconstruct their track if a pixelated anode is used. In our case, the gasmore » is used as a (n, p) converter in order to detect neutrons down to few keV. Coming from elastic collisions with neutrons, recoil protons lose a part of their kinetic energy by ionizing the gas. The ionization electrons are drifted toward a pixelated anode (2D projection), read at 50 MHz by a self-triggered electronic system to obtain the third track dimension. The neutron energy is reconstructed event by event thanks to proton scattering angle and proton energy measurements. The scattering angle is deduced from the 3D track. The proton energy is obtained by charge collection measurements, knowing the ionization quenching factor (i.e. the part of proton kinetic energy lost by ionizing the gas). The fluence is calculated thanks to the detected events number and the simulation of the detector response. The μTPC is a new reliable detector able to measure energy distribution of the neutron fluence without unfolding procedure or prior neutron calibration contrary to usual gaseous counters. The microTPC is still being developed and measurements have been carried out at the AMANDE facility, with neutrons energies going from 8 keV to 565 keV. After the context and the μ-TPC working principle presentation, measurements of the neutron energy and fluence at 27 keV and 144 keV are shown and compared to the complete detector response simulation. This

  1. Electron currents associated with an auroral band

    NASA Technical Reports Server (NTRS)

    Spiger, R. J.; Anderson, H. R.

    1975-01-01

    Measurements of electron pitch angle distributions and energy spectra over a broad auroral band were used to calculate net electric current carried by auroral electrons in the vicinity of the band. The particle energy 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 energy 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.

  2. Sub-arcsec X-Ray Telescope for Imaging The Solar Corona In the 0.25 - 1.2 keV Band

    NASA Technical Reports Server (NTRS)

    Gallagher, Dennis; Cash, Webster; Jelsma, Schuyler; Farmer, Jason

    1996-01-01

    We have developed an X-ray telescope that uses a new technique for focusing X-rays with grazing incidence optics. The telescope was built with spherical optics for all of its components, utilizing the high quality surfaces obtainable when polishing spherical (as opposed to aspherical) optics. We tested the prototype X-ray telescope in the 300 meter vacuum pipe at White Sands Missile Range, NM. The telescope features 2 degee graze angles with tungsten coatings, yielding a bandpass of 0.25-1.5 keV with a peak effective area of 0.8 sq cm at 0.83 keV. Results from X-ray testing at energies of 0.25 keV and 0.93 keV (C-K and Cu-L) verify 0.5 arcsecond performance at 0.93 keV. Results from modeling the X-ray telescope's response to the Sun show that the current design would be capable of recording 10 half arcsecond images of a solar active region during a 300 second NASA sounding rocket flight.

  3. DISPLACEMENT CASCADE SIMULATION IN TUNGSTEN UP TO 200 KEV OF DAMAGE ENERGY AT 300, 1025, AND 2050 K

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    Setyawan, Wahyu; Nandipati, Giridhar; Roche, Kenneth J.

    2015-09-22

    We generated molecular dynamics database of primary defects that adequately covers the range of tungsten recoil energy imparted by 14-MeV neutrons. During this semi annual period, cascades at 150 and 200 keV at 300 and 1025 K were simulated. Overall, we included damage energy up to 200 keV at 300 and 1025 K, and up to 100 keV at 2050 K. We report the number of surviving Frenkel pairs (NF) and the size distribution of defect clusters. The slope of the NF curve versus cascade damage energy (EMD), on a log-log scale, changes at a transition energy (μ). For EMDmore » > μ, the cascade forms interconnected damage regions that facilitate the formation of large clusters of defects. At 300 K and EMD = 200 keV, the largest size of interstitial cluster and vacancy cluster is 266 and 335, respectively. Similarly, at 1025 K and EMD = 200 keV, the largest size of interstitial cluster and vacancy cluster is 296 and 338, respectively. At 2050 K, large interstitial clusters also routinely form, but practically no large vacancy clusters do« less

  4. Energy staggering in superdeformed bands in {sup 131}Ce, {sup 132}Ce, and {sup 133}Ce

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    Semple, A.T.; Nolan, P.J.; Beausang, C.W.

    1996-05-01

    Superdeformed bands observed in {sup 131}Ce, {sup 132}Ce, and {sup 133}Ce have sequences of {gamma}-ray transition energies 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 energy 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.}

  5. Calibration of hard x-ray (15 - 50 keV) optics at the MPE test facility PANTER

    NASA Astrophysics Data System (ADS)

    Bräuninger, Heinrich; Burkert, Wolfgang; Hartner, Gisela D.; Citterio, Oberto; Ghigo, Mauro; Mazzoleni, Francesco; Pareschi, Giovanni; Spiga, Daniele

    2004-02-01

    The Max-Planck-Institut für extraterrestrische Physik (MPE) in Garching, Germany, operates the large X-ray beam line facility PANTER for testing astronomical systems. At PANTER a number of telescopes like EXOSAT, ROSAT, SAX, JET-X, ABRIXAS, XMM and SWIFT operating in the soft energy range (0.02 - 15 keV) have been successfully calibrated. In the present paper we report on an important upgrade recently implemented that enables the calibration of hard X-ray optics (from 15 up to 50 keV). Currently hard X-ray optics based on single and multilayer coating are being developed for several future X-ray missions. The hard X-ray calibrations at PANTER are carried out by a high energy source based on an electron gun and several anodes, able to cover the energy range from 4.5 up to 50 keV. It provides fluxes up to 104 counts/sec/cm2 at the instrument chamber with a stability better than 1%. As detector a pn-CCD camera operating between 0.2 and 50 keV and a collecting area of 36 cm2 is used. Taking into account the high energy resolution of the CCD (145 eV at 6 keV), a very easy way to operate the facility in hard X-ray is in energy-dispersive mode (i.e. with a broad-band beam). A double crystal monochromator is also available providing energies up to 20 keV. In this paper we present the first results obtained by using PANTER for hard X-ray characterizations, performed on prototype multilayer optics developed by the Osservatorio Astronomico di Brera (OAB), Milano, Italy, and the Harvard-Smithsonian Center for Astrophysics (CfA), Cambridge, MA, USA.

  6. A balloon-borne instrument for high-resolution astrophysical spectroscopy in the 20-8000 keV energy range

    NASA Technical Reports Server (NTRS)

    Paciesas, W. S.; Baker, R.; Boclet, D.; Brown, S.; Cline, T.; Costlow, H.; Durouchoux, P.; Ehrmann, C.; Gehrels, N.; Hameury, J. M.

    1983-01-01

    The Low Energy Gamma ray Spectrometer (LEGS) is designed to perform fine energy resolution measurements of astrophysical sources. The instrument is configured for a particular balloon flight with either of two sets of high purity germanium detectors. In one configuration, the instrument uses an array of three coaxial detectors (effective volume equal to or approximately 230 cubic cm) inside an NaI (T1) shield and collimator (field of view equal to or approximately 16 deg FWHM) and operates in the 80 to 8000 keV energy range. In the other configuration, three planar detectors (effective area equal to or approximately square cm) surrounded by a combination of passive Fe and active NaI for shielding and collimation (field of view equal to or approximately 5 deg x 10 deg FWHM) are optimized for the 20 to 200 keV energy range. In a typical one day balloon flight, LEGS sensitivity limit (3 sigma) for narrow line features is less than or approximately .0008 ph/cm/s square (coaxial array: 80 to 2000 keV) and less than or approximately .0003 ph/square cm/s (planar array: 50 to 150 keV).

  7. Applications of “Tender” Energy (1-5 keV) X-ray Absorption Spectroscopy in Life Sciences

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    Northrup, Paul; Leri, Alessandra; Tappero, Ryan

    The “tender” energy range of 1 to 5 keV, between the energy ranges of most “hard” (>5 keV) and “soft” (<1 keV) synchrotron X-ray facilities, offers some unique opportunities for synchrotron-based X-ray absorption fine structure spectroscopy in life sciences. In particular the K absorption edges of Na through Ca offer opportunities to study local structure, speciation, and chemistry of many important biological compounds, structures and processes. This is an area of largely untapped science, in part due to a scarcity of optimized facilities. Such measurements also entail unique experimental challenges. Lastly, this brief review describes the technique, its experimental challenges,more » recent progress in development of microbeam measurement capabilities, and several highlights illustrating applications in life sciences.« less

  8. Applications of “Tender” Energy (1-5 keV) X-ray Absorption Spectroscopy in Life Sciences

    DOE PAGES

    Northrup, Paul; Leri, Alessandra; Tappero, Ryan

    2016-02-15

    The “tender” energy range of 1 to 5 keV, between the energy ranges of most “hard” (>5 keV) and “soft” (<1 keV) synchrotron X-ray facilities, offers some unique opportunities for synchrotron-based X-ray absorption fine structure spectroscopy in life sciences. In particular the K absorption edges of Na through Ca offer opportunities to study local structure, speciation, and chemistry of many important biological compounds, structures and processes. This is an area of largely untapped science, in part due to a scarcity of optimized facilities. Such measurements also entail unique experimental challenges. Lastly, this brief review describes the technique, its experimental challenges,more » recent progress in development of microbeam measurement capabilities, and several highlights illustrating applications in life sciences.« less

  9. ART: Surveying the Local Universe at 2-11 keV

    NASA Technical Reports Server (NTRS)

    O'Dell, S. L.; Ramsey, B. D.; Adams, M. L.; Brandt, W. N.; Bubarev, M. V.; Hassinger, G.; Pravlinski, M.; Predehl, P.; Romaine, S. E.; Swartz, D. A.; hide

    2008-01-01

    The Astronomical Rontgen Telescope (ART) is a medium-energy x-ray telescope system proposed for the Russian-led mission Spectrum Rontgen-Gamma (SRG). Optimized for performance over the 2-11-keV band, ART complements the softer response of the SRG prime instrument-the German eROSITA x-ray telescope system. The anticipated number of ART detections is 50,000-with 1,000 heavily-obscured (N(sub H)> 3x10(exp 23)/sq cm) AGN-in the SRG 4-year all-sky survey, plus a comparable number in deeper wide-field (500 deg(sup 2) total) surveys. ART's surveys will provide a minimally-biased, nearly-complete census of the local Universe in the medium-energy x-ray band (including Fe-K lines), at CCD spectral resolution. During long (approx.100-ks) pointed observations, ART can obtain statistically significant spectral data up to about 15 keY for bright sources and medium-energy x-ray continuum and Fe-K-line spectra of AGN detected with the contemporaneous NuSTAR hard-x-ray mission.

  10. A comparative study of inelastic scattering models at energy levels ranging from 0.5 keV to 10 keV

    NASA Astrophysics Data System (ADS)

    Hu, Chia-Yu; Lin, Chun-Hung

    2017-03-01

    Six models, including a single-scattering model, four hybrid models, and one dielectric function model, were evaluated using Monte Carlo simulations for aluminum and copper at incident beam energies ranging from 0.5 keV to 10 keV. The inelastic mean free path, mean energy loss per unit path length, and backscattering coefficients obtained by these models are compared and discussed to understand the merits of the various models. ANOVA (analysis of variance) statistical models were used to quantify the effects of inelastic cross section and energy loss models on the basis of the simulated results deviation from the experimental data for the inelastic mean free path, the mean energy loss per unit path length, and the backscattering coefficient, as well as their correlations. This work in this study is believed to be the first application of ANOVA models towards evaluating inelastic electron beam scattering models. This approach is an improvement over the traditional approach which involves only visual estimation of the difference between the experimental data and simulated results. The data suggests that the optimization of the effective electron number per atom, binding energy, and cut-off energy of an inelastic model for different materials at different beam energies is more important than the selection of inelastic models for Monte Carlo electron scattering simulation. During the simulations, parameters in the equations should be tuned according to different materials for different beam energies rather than merely employing default parameters for an arbitrary material. Energy loss models and cross-section formulas are not the main factors influencing energy loss. Comparison of the deviation of the simulated results from the experimental data shows a significant correlation (p < 0.05) between the backscattering coefficient and energy loss per unit path length. The inclusion of backscattering electrons generated by both primary and secondary electrons for

  11. Surface damage studies of ETFE polymer bombarded with low energy Si ions (⩽100 keV)

    NASA Astrophysics Data System (ADS)

    Minamisawa, Renato Amaral; Almeida, Adelaide De; Budak, Satilmis; Abidzina, Volha; Ila, Daryush

    2007-08-01

    Surface studies of ethylenetetrafluoroethylene (ETFE), bombarded with Si in a high-energy tandem Pelletron accelerator, have recently been reported. Si ion bombardment with a few MeV to a few hundred keV energies was shown to be sufficient to produce damage on ETFE film. We report here the use of a low energy implanter with Si ion energies lower than 100 keV, to induce changes on ETFE films. In order to determine the radiation damage, ETFE bombarded films were simulated with SRIM software and analyzed with optical absorption photometry (OAP), Raman and Fourier transform infrared-attenuated total reflectance (FTIR-ATR) spectroscopy to show quantitatively the physical and chemical property changes. Carbonization occurs following higher dose implantation, and hydroperoxides were formed following dehydroflorination of the polymer.

  12. Studies on effective atomic numbers for photon energy absorption and electron density of some narcotic drugs in the energy range 1 keV-20 MeV

    NASA Astrophysics Data System (ADS)

    Gounhalli, Shivraj G.; Shantappa, Anil; Hanagodimath, S. M.

    2013-04-01

    Effective atomic numbers for photon energy absorption ZPEA,eff, photon interaction ZPI,eff and for electron density Nel, have been calculated by a direct method in the photon-energy region from 1 keV to 20 MeV for narcotic drugs, such as Heroin (H), Cocaine (CO), Caffeine (CA), Tetrahydrocannabinol (THC), Cannabinol (CBD), Tetrahydrocannabivarin (THCV). The ZPEA,eff, ZPI,eff and Nel values have been found to change with energy and composition of the narcotic drugs. The energy dependence ZPEA,eff, ZPI,eff and Nel is shown graphically. The maximum difference between the values of ZPEA,eff, and ZPI,eff occurs at 30 keV and the significant difference of 2 to 33% for the energy region 5-100 keV for all drugs. The reason for these differences is discussed.

  13. Microsecond Electron Beam Source with Electron Energy Up to 400 Kev and Plasma Anode

    NASA Astrophysics Data System (ADS)

    Abdullin, É. N.; Basov, G. F.; Shershnev, S.

    2017-12-01

    A new high-power source of electrons with plasma anode for producing high-current microsecond electron beams with electron energy up to 400 keV has been developed, manufactured, and put in operation. To increase the cross section and pulse current duration of the beam, a multipoint explosive emission cathode is used in the electron beam source, and the beam is formed in an applied external guiding magnetic field. The Marx generator with vacuum insulation is used as a high-voltage source. Electron beams with electron energy up to 300-400 keV, current of 5-15 kA, duration of 1.5-3 μs, energy up to 4 kJ, and cross section up to 150 cm2 have been produced. The operating modes of the electron beam source are realized in which the applied voltage is influenced weakly on the current. The possibility of source application for melting of metal surfaces is demonstrated.

  14. 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('https://www.ncbi.nlm.nih.gov/pubmed/21928623','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21928623"><span>On the Monte Carlo simulation of electron transport in the sub-1 <span class="hlt">keV</span> <span class="hlt">energy</span> range.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Thomson, Rowan M; Kawrakow, Iwan</p> <p>2011-08-01</p> <p>The validity of "classic" Monte Carlo (MC) simulations of electron and positron transport at sub-1 <span class="hlt">keV</span> <span class="hlt">energies</span> is investigated in the context of quantum theory. Quantum theory dictates that uncertainties on the position and <span class="hlt">energy</span>-momentum four-vectors of radiation quanta obey Heisenberg's uncertainty relation; however, these uncertainties are neglected in "classical" MC simulations of radiation transport in which position and momentum are known precisely. Using the quantum uncertainty relation and electron mean free path, the magnitudes of uncertainties on electron position and momentum are calculated for different kinetic <span class="hlt">energies</span>; a validity bound on the classical simulation of electron transport is derived. In order to satisfy the Heisenberg uncertainty principle, uncertainties of 5% must be assigned to position and momentum for 1 <span class="hlt">keV</span> electrons in water; at 100 eV, these uncertainties are 17 to 20% and are even larger at lower <span class="hlt">energies</span>. In gaseous media such as air, these uncertainties are much smaller (less than 1% for electrons with <span class="hlt">energy</span> 20 eV or greater). The classical Monte Carlo transport treatment is questionable for sub-1 <span class="hlt">keV</span> electrons in condensed water as uncertainties on position and momentum must be large (relative to electron momentum and mean free path) to satisfy the quantum uncertainty principle. Simulations which do not account for these uncertainties are not faithful representations of the physical processes, calling into question the results of MC track structure codes simulating sub-1 <span class="hlt">keV</span> electron transport. Further, the large difference in the scale at which quantum effects are important in gaseous and condensed media suggests that track structure measurements in gases are not necessarily representative of track structure in condensed materials on a micrometer or a nanometer scale.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017ResPh...7..272A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017ResPh...7..272A"><span>Gamma ray interaction studies of organic nonlinear optical materials in the <span class="hlt">energy</span> range 122 <span class="hlt">keV</span>-1330 <span class="hlt">keV</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Awasarmol, V. V.; Gaikwad, D. K.; Raut, S. D.; Pawar, P. P.</p> <p></p> <p>The mass attenuation coefficients (μm) for organic nonlinear optical materials measured at 122-1330 <span class="hlt">keV</span> photon <span class="hlt">energies</span> were investigated on the basis of mixture rule and compared with obtained values of WinXCOM program. It is observed that there is a good agreement between theoretical and experimental values of the samples. All samples were irradiated with six radioactive sources such as 57Co, 133Ba, 22Na, 137Cs, 54Mn and 60Co using transmission arrangement. Effective atomic and electron numbers or electron densities (Zeff and Neff), molar extinction coefficient (ε), mass <span class="hlt">energy</span> absorption coefficient (μen/ρ) and effective atomic <span class="hlt">energy</span> absorption cross section (σa,en) were determined experimentally and theoretically using the obtained μm values for investigated samples and graphs have been plotted. The graph shows that the variation of all samples decreases with increasing photon <span class="hlt">energy</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26395374','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26395374"><span>Iodine-131 imaging using 284 <span class="hlt">keV</span> photons with a small animal CZT-SPECT system dedicated to low-medium-<span class="hlt">energy</span> photon detection.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kojima, Akihiro; Gotoh, Kumiko; Shimamoto, Masako; Hasegawa, Koki; Okada, Seiji</p> <p>2016-02-01</p> <p>Iodine-131 is widely used for radionuclide therapy because of its β-particle and for diagnostic imaging employing its principal gamma ray. Since that principal gamma ray has the relatively high <span class="hlt">energy</span> of 364 <span class="hlt">keV</span>, small animal single-photon emission computed tomography (SPECT) imaging systems may be required to possess the ability to image such higher <span class="hlt">energy</span> photons. The aim of this study was to investigate the possibility of imaging I-131 using its 284 <span class="hlt">keV</span> photons instead of its 364 <span class="hlt">keV</span> photons in a small animal SPECT imaging system dedicated to the detection of low-medium-<span class="hlt">energy</span> photons (below 300 <span class="hlt">keV</span>). The imaging system used was a commercially available preclinical SPECT instrument with CZT detectors that was equipped with multi-pinhole collimators and was accompanied by a CT imager. An <span class="hlt">energy</span> window for I-131 imaging was set to a photopeak of 284 <span class="hlt">keV</span> with a low abundance compared with 364 <span class="hlt">keV</span> photons. Small line sources and two mice, one of each of two types, that were injected with NaI-131 were scanned. Although higher counts occurred at the peripheral region of the reconstructed images due to the collimator penetration by the 364 <span class="hlt">keV</span> photons, the shape of the small line sources could be well visualized. The measured spatial resolution was relatively poor (~1.9 mm for full width at half maximum and ~3.9 mm for full width at tenth maximum). However, a good linear correlation between SPECT values and the level of I-131 radioactivity was observed. Furthermore, the uptake of NaI-131 to the thyroid gland for the two mice was clearly identified in the 3D-SPECT image fused with the X-ray CT image. We conclude that the use of an <span class="hlt">energy</span> window set on the photopeak of 284 <span class="hlt">keV</span> and the multi-pinhole collimator may permit I-131 imaging for a preclinical CZT-SPECT system that does not have the ability to acquire images using the 364 <span class="hlt">keV</span> photons.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19930037267&hterms=Day+George&qs=N%3D0%26Ntk%3DAuthor-Name%26Ntx%3Dmode%2Bmatchall%26Ntt%3DDay%252C%2BGeorge%2BS','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19930037267&hterms=Day+George&qs=N%3D0%26Ntk%3DAuthor-Name%26Ntx%3Dmode%2Bmatchall%26Ntt%3DDay%252C%2BGeorge%2BS"><span>The broad-<span class="hlt">band</span> X-ray spectral variability of Mrk 841</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>George, I. M.; Nandra, K.; Fabian, A. C.; Turner, T. J.; Done, C.; Day, C. S. R.</p> <p>1993-01-01</p> <p>A detailed spectral analysis of five X-ray observations of Mrk 841 with the EXOSAT, Ginga, and ROSAT satellites is reported. Variability is apparent in both the soft (0.1-1.0 <span class="hlt">keV</span>) and medium (1-20 <span class="hlt">keV</span>) <span class="hlt">energy</span> <span class="hlt">bands</span>. Above, 1 <span class="hlt">keV</span>, the spectra are adequately modeled by a power law with a strong emission line of equivalent width 450 eV. The large equivalent width of the emission line indicates a strongly enhanced reflection component of the source compared with other Seyferts observed with Ginga. The implications of the results of the analysis for physical models of the emission regions in this and other X-ray bright Seyferts are briefly examined.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017GeoRL..4411722W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017GeoRL..4411722W"><span>The Effects of Hydrogen <span class="hlt">Band</span> EMIC Waves on Ring Current H+ Ions</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; Zhai, Hao; Gao, Zhuxiu</p> <p>2017-12-01</p> <p>Hydrogen <span class="hlt">band</span> electromagnetic ion cyclotron (EMIC) waves have received much attention recently because they are found to frequently span larger spatial areas than the other <span class="hlt">band</span> EMIC waves. Using test particle simulations, we study the nonlinear effects of hydrogen <span class="hlt">band</span> EMIC waves on ring current H+ ions. A dimensionless parameter R is used to characterize the competition between wave-induced and adiabatic motions. The results indicate that there are three regimes of wave-particle interactions for typical 35 <span class="hlt">keV</span> H+ ions at L = 5: diffusive (quasi-linear) behavior when αeq ≤ 35° (R ≥ 2.45), the nonlinear phase trapping when 35° < αeq < 50° (0.75 < R < 2.45), and both the nonlinear phase bunching and phase trapping when αeq ≥ 50° (R ≤ 0.75). The phase trapping can transport H+ ions toward large pitch angle, while the phase bunching has the opposite effect. The phase-trapped H+ ions can be significantly accelerated (from 35 <span class="hlt">keV</span> to over 500 <span class="hlt">keV</span>) in about 4 min and thus contribute to the formation of high <span class="hlt">energy</span> components of ring current ions. The results suggest that the effect of hydrogen <span class="hlt">band</span> EMIC waves is not ignorable in the nonlinear acceleration and resonance scattering of ring current H+ ions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2004AIPC..705.1328T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2004AIPC..705.1328T"><span>Interferometric phase-contrast X-ray CT imaging of VX2 rabbit cancer at 35<span class="hlt">keV</span> X-ray <span class="hlt">energy</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Takeda, Tohoru; Wu, Jin; Tsuchiya, Yoshinori; Yoneyama, Akio; Lwin, Thet-Thet; Hyodo, Kazuyuki; Itai, Yuji</p> <p>2004-05-01</p> <p>Imaging of large objects at 17.7-<span class="hlt">keV</span> low x-ray <span class="hlt">energy</span> causes huge x-ray exposure to the objects even using interferometric phase-contrast x-ray CT (PCCT). Thus, we tried to obtain PCCT images at high x-ray <span class="hlt">energy</span> of 35<span class="hlt">keV</span> and examined the image quality using a formalin-fixed VX2 rabbit cancer specimen with 15-mm in diameter. The PCCT system consisted of an asymmetrically cut silicon (220) crystal, a monolithic x-ray interferometer, a phase-shifter, an object cell and an x-ray CCD camera. The PCCT at 35 <span class="hlt">keV</span> clearly visualized various inner structures of VX2 rabbit cancer such as necrosis, cancer, the surrounding tumor vessels, and normal liver tissue. Besides, image-contrast was not degraded significantly. These results suggest that the PCCT at 35 <span class="hlt">KeV</span> is sufficient to clearly depict the histopathological morphology of VX2 rabbit cancer specimen.</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/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 <span class="hlt">keV</span> 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/2018ResPh...8..683N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018ResPh...8..683N"><span>Slowing down of 100 <span class="hlt">keV</span> antiprotons in Al foils</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Nordlund, K.</p> <p>2018-03-01</p> <p>Using <span class="hlt">energy</span> degrading foils to slow down antiprotons is of interest for producing antihydrogen atoms. I consider here the slowing down of 100 <span class="hlt">keV</span> antiprotons, that will be produced in the ELENA storage ring under construction at CERN, to <span class="hlt">energies</span> below 10 <span class="hlt">keV</span>. At these low <span class="hlt">energies</span>, they are suitable for efficient antihydrogen production. I simulate the antihydrogen motion and slowing down in Al foils using a recently developed molecular dynamics approach. The results show that the optimal Al foil thickness for slowing down the antiprotons to below 5 <span class="hlt">keV</span> is 910 nm, and to below 10 <span class="hlt">keV</span> is 840 nm. Also the lateral spreading of the transmitted antiprotons is reported and the uncertainties discussed.</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 <span class="hlt">keV</span>) 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/2016A%26A...587A.142F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016A%26A...587A.142F"><span>The 5-10 <span class="hlt">keV</span> AGN luminosity function at 0.01 < z < 4.0</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Fotopoulou, S.; Buchner, J.; Georgantopoulos, I.; Hasinger, G.; Salvato, M.; Georgakakis, A.; Cappelluti, N.; Ranalli, P.; Hsu, L. T.; Brusa, M.; Comastri, A.; Miyaji, T.; Nandra, K.; Aird, J.; Paltani, S.</p> <p>2016-03-01</p> <p>The active galactic nuclei (AGN) X-ray luminosity function traces actively accreting supermassive black holes and is essential for the study of the properties of the AGN population, black hole evolution, and galaxy-black hole coevolution. Up to now, the AGN luminosity function has been estimated several times in soft (0.5-2 <span class="hlt">keV</span>) and hard X-rays (2-10 <span class="hlt">keV</span>). AGN selection in these <span class="hlt">energy</span> ranges often suffers from identification and redshift incompleteness and, at the same time, photoelectric absorption can obscure a significant amount of the X-ray radiation. We estimate the evolution of the luminosity function in the 5-10 <span class="hlt">keV</span> <span class="hlt">band</span>, where we effectively avoid the absorbed part of the spectrum, rendering absorption corrections unnecessary up to NH ~ 1023 cm-2. Our dataset is a compilation of six wide, and deep fields: MAXI, HBSS, XMM-COSMOS, Lockman Hole, XMM-CDFS, AEGIS-XD, Chandra-COSMOS, and Chandra-CDFS. This extensive sample of ~1110 AGN (0.01 < z < 4.0, 41 < log Lx < 46) is 98% redshift complete with 68% spectroscopic redshifts. For sources lacking a spectroscopic redshift estimation we use the probability distribution function of photometric redshift estimation specifically tuned for AGN, and a flat probability distribution function for sources with no redshift information. We use Bayesian analysis to select the best parametric model from simple pure luminosity and pure density evolution to more complicated luminosity and density evolution and luminosity-dependent density evolution (LDDE). We estimate the model parameters that describe best our dataset separately for each survey and for the combined sample. We show that, according to Bayesian model selection, the preferred model for our dataset is the LDDE. Our estimation of the AGN luminosity function does not require any assumption on the AGN absorption and is in good agreement with previous works in the 2-10 <span class="hlt">keV</span> <span class="hlt">energy</span> <span class="hlt">band</span> based on X-ray hardness ratios to model the absorption in AGN up to redshift three</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2006SPIE.6266E..2AC','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2006SPIE.6266E..2AC"><span>A focal plane detector design for a wide <span class="hlt">band</span> Laue-lens telescope</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Caroli, E.; Auricchio, N.; Bertuccio, G.; Budtz-Jørgensen, C.; Curado da Silva, R. M.; Del Sordo, S.; Frontera, F.; Quadrini, E.; Ubertini, P.; Ventura, G.</p> <p>2006-06-01</p> <p>The <span class="hlt">energy</span> range above 50 <span class="hlt">keV</span> is important for the study of many open problems in high <span class="hlt">energy</span> astrophysics such as, non thermal mechanisms in SNR, the study of the high <span class="hlt">energy</span> cut-offs in AGN spectra, and the detection of nuclear and annihilation lines. In the framework of the definition of a new mission concept for hard X and soft gamma ray (GRI- Gamma Ray Imager) for the next decade, the use of Laue lenses with broad <span class="hlt">energy</span> <span class="hlt">band</span>-passes from 100 to 1000 <span class="hlt">keV</span> is under study. This kind of instruments will be used for deep study the hard X-ray continuum of celestial sources. This new telescope will require focal plane detectors with high detection efficiency over the entire operative range, an <span class="hlt">energy</span> resolution of few <span class="hlt">keV</span> at 500 <span class="hlt">keV</span> and a sensitivity to linear polarization. We describe a possible configuration for the focal plane detector based on CdTe/CZT pixelated layers stacked together to achieve the required detection efficiency at high <span class="hlt">energy</span>. Each layer can either operate as a separate position sensitive detector and a polarimeter or together with other layers in order to increase the overall full <span class="hlt">energy</span> efficiency. We report on the current state of art in high Z spectrometers development and on some activities undergoing. Furthermore we describe the proposed focal plane option with the required resources and an analytical summary of the achievable performance in terms of efficiency and polarimetry.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012AGUFMSH33A2219C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012AGUFMSH33A2219C"><span>A New Observation of the Quiet Sun Soft X-ray (0.5-5 <span class="hlt">keV</span>) Spectrum</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Caspi, A.; Woods, T. N.; Stone, J.</p> <p>2012-12-01</p> <p>The solar corona is the brightest source of X-rays in the solar system, and the X-ray emission is highly variable with solar activity. While this is particularly true during solar flares, when emission can be enhanced by many orders of magnitude up to gamma-ray <span class="hlt">energies</span>, even the so-called "quiet Sun" is bright in soft X-rays (SXRs), as the ~1-2 MK ambient plasma of the corona emits significant thermal bremsstrahlung up to ~5 <span class="hlt">keV</span>. However, the actual solar SXR (0.5-5 <span class="hlt">keV</span>) spectrum is not well known, particularly during quiet periods, as, with few exceptions, this <span class="hlt">energy</span> range has not been systematically studied in many years. Previous observations include ultra-high-resolution but very narrow-<span class="hlt">band</span> spectra from crystral spectrometers (e.g. Yohkoh/BCS), or integrated broadband irradiances from photometers (e.g. GOES/XRS, TIMED/XPS, etc.) that lack detailed spectral information. In recent years, broadband measurements with fair <span class="hlt">energy</span> resolution (~0.5-0.7 <span class="hlt">keV</span> FWHM) were made by SphinX on CORONAS-Photon and XRS on MESSENGER, although they did not extend below ~1 <span class="hlt">keV</span>. We present observations of the quiet Sun SXR emission obtained using a new SXR spectrometer flown on the third SDO/EVE underflight calibration rocket (NASA 36.286). The commercial off-the-shelf Amptek X123 silicon drift detector, with an 8-micron Be window and custom aperture, measured the solar SXR emission from ~0.5 to >10 <span class="hlt">keV</span> with ~0.15 <span class="hlt">keV</span> FWHM resolution (though, due to hardware limitations, with only ~0.12 <span class="hlt">keV</span> binning) and 2-sec cadence over ~5 minutes on 23 June 2012. Despite the rising solar cycle, activity on 23 June 2012 was abnormally low, with no visible active regions and GOES XRS emission near 2010 levels; we measured no solar counts above ~4 <span class="hlt">keV</span> during the observation period. We compare our X123 measurements with spectra and broadband irradiances from other instruments, including the SphinX observations during the deep solar minimum of 2009, and with upper limits of >3 <span class="hlt">keV</span> quiet Sun emission</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013enss.confE..73C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013enss.confE..73C"><span>A New Observation of the Quiet Sun Soft X-ray (0.5-5 <span class="hlt">keV</span>) Spectrum</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Caspi, Amir; Woods, Thomas N.; Stone, Jordan</p> <p>2013-03-01</p> <p>The solar corona is the brightest source of X-rays in the solar system, and the X-ray emission is highly variable with solar activity. While this is particularly true during solar flares, when emission can be enhanced by many orders of magnitude up to gamma-ray <span class="hlt">energies</span>, even the so-called "quiet Sun" is bright in soft X-rays (SXRs), as the 1-2 MK ambient plasma of the corona emits significant thermal bremsstrahlung up to 5 <span class="hlt">keV</span>. However, the actual solar SXR (0.5-5 <span class="hlt">keV</span>) spectrum is not well known, particularly during quiet periods, as, with few exceptions, this <span class="hlt">energy</span> range has not been systematically studied in many years. Previous observations include ultra-high-resolution but very narrow-<span class="hlt">band</span> spectra from crystral spectrometers (e.g. Yohkoh/BCS), or integrated broadband irradiances from photometers (e.g. GOES/XRS, TIMED/XPS, etc.) that lack detailed spectral information. In recent years, broadband measurements with fair <span class="hlt">energy</span> resolution ( 0.5-0.7 <span class="hlt">keV</span> FWHM) were made by SphinX on CORONAS-Photon and XRS on MESSENGER, although they did not extend below 1 <span class="hlt">keV</span>. We present observations of the quiet Sun SXR emission obtained using a new SXR spectrometer flown on the third SDO/EVE underflight calibration rocket (NASA 36.286). The commercial off-the-shelf Amptek X123 silicon drift detector, with an 8-micron Be window and custom aperture, measured the solar SXR emission from 0.5 to >10 <span class="hlt">keV</span> with 0.15 <span class="hlt">keV</span> FWHM resolution (though, due to hardware limitations, with only 0.12 <span class="hlt">keV</span> binning) and 2-sec cadence over 5 minutes on 23 June 2012. Despite the rising solar cycle, activity on 23 June 2012 was abnormally low, with no visible active regions and GOES XRS emission near 2010 levels; we measured no solar counts above 4 <span class="hlt">keV</span> during the observation period. We compare our X123 measurements with spectra and broadband irradiances from other instruments, including the SphinX observations during the deep solar minimum of 2009, and with upper limits of >3 <span class="hlt">keV</span> quiet Sun emission</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013AGUFMSH33A2036C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013AGUFMSH33A2036C"><span>New Observations of Soft X-ray (0.5-5 <span class="hlt">keV</span>) Solar Spectra</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Caspi, A.; Woods, T. N.; Mason, J. P.; Jones, A. R.; Warren, H. P.</p> <p>2013-12-01</p> <p>The solar corona is the brightest source of X-rays in the solar system, and the X-ray emission is highly variable on many time scales. However, the actual solar soft X-ray (SXR) (0.5-5 <span class="hlt">keV</span>) spectrum is not well known, particularly during solar quiet periods, as, with few exceptions, this <span class="hlt">energy</span> range has not been systematically studied in many years. Previous observations include high-resolution but very narrow-<span class="hlt">band</span> spectra from crystal spectrometers (e.g., Yohkoh/BCS), or integrated broadband irradiances from photometers (e.g., GOES/XRS, TIMED/XPS, etc.) that lack detailed spectral information. In recent years, broadband measurements with moderate <span class="hlt">energy</span> resolution (~0.5-0.7 <span class="hlt">keV</span> FWHM) were made by SphinX on CORONAS-Photon and SAX on MESSENGER, although they did not extend to <span class="hlt">energies</span> below ~1 <span class="hlt">keV</span>. We present observations of solar SXR emission obtained using new instrumentation flown on recent SDO/EVE calibration rocket underflights. The photon-counting spectrometer, a commercial Amptek X123 with a silicon drift detector and an 8 μm Be window, measures the solar disk-integrated SXR emission from ~0.5 to >10 <span class="hlt">keV</span> with ~0.15 <span class="hlt">keV</span> FWHM resolution and 1 s cadence. A novel imager, a pinhole X-ray camera using a cooled frame-transfer CCD (15 μm pixel pitch), Ti/Al/C filter, and 5000 line/mm Au transmission grating, images the full Sun in multiple spectral orders from ~0.1 to ~5 nm with ~10 arcsec/pixel and ~0.01 nm/pixel spatial and spectral detector scales, respectively, and 10 s cadence. These instruments are prototypes for future CubeSat missions currently being developed. We present new results of solar observations on 04 October 2013 (NASA sounding rocket 36.290). We compare with previous results from 23 June 2012 (NASA sounding rocket 36.286), during which solar activity was low and no signal was observed above ~4 <span class="hlt">keV</span>. We compare our spectral and imaging measurements with spectra and broadband irradiances from other instruments, including SDO/EVE, GOES/XRS, TIMED</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018PhRvB..97m4104R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PhRvB..97m4104R"><span><span class="hlt">Energy</span> dependence of the spatial distribution of inelastically scattered electrons in backscatter electron diffraction</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ram, Farangis; De Graef, Marc</p> <p>2018-04-01</p> <p>In an electron backscatter diffraction pattern (EBSP), the angular distribution of backscattered electrons (BSEs) depends on their <span class="hlt">energy</span>. Monte Carlo modeling of their depth and <span class="hlt">energy</span> distributions suggests that the highest <span class="hlt">energy</span> BSEs are more likely to hit the bottom of the detector than the top. In this paper, we examine experimental EBSPs to validate the modeled angular BSE distribution. To that end, the Kikuchi bandlet method is employed to measure the width of Kikuchi <span class="hlt">bands</span> in both modeled and measured EBSPs. The results show that in an EBSP obtained with a 15 <span class="hlt">keV</span> primary probe, the width of a Kikuchi <span class="hlt">band</span> varies by about 0 .4∘ from the bottom of the EBSD detector to its top. The same is true for a simulated pattern that is composed of BSEs with 5 <span class="hlt">keV</span> to 15 <span class="hlt">keV</span> <span class="hlt">energies</span>, which validates the Monte Carlo simulations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20170000982&hterms=1089&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3D%2526%25231089','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20170000982&hterms=1089&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3D%2526%25231089"><span><span class="hlt">Banded</span> Structures in Electron Pitch Angle Diffusion Coefficients from Resonant Wave Particle Interactions</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Tripathi, A. K.; Singhal, R. P.; Khazanov, G. V.; Avanov, L. A.</p> <p>2016-01-01</p> <p>Electron pitch angle (D (alpha)) and momentum (D(pp)) diffusion coefficients have been calculated due to resonant interactions with electrostatic electron cyclotron harmonic (ECH) and whistler mode chorus waves. Calculations have been performed at two spatial locations L = 4.6 and 6.8 for electron <span class="hlt">energies</span> 10 <span class="hlt">keV</span>. Landau (n = 0) resonance and cyclotron harmonic resonances n = +/-1, +/-2,...+/-5 have been included in the calculations. It is found that diffusion coefficient versus pitch angle (alpha) profiles show large dips and oscillations or <span class="hlt">banded</span> structures. The structures are more pronounced for ECH and lower <span class="hlt">band</span> chorus (LBC) and particularly at location 4.6. Calculations of diffusion coefficients have also been performed for individual resonances. It is noticed that the main contribution of ECH waves in pitch angle diffusion coefficient is due to resonances n = +1 and n = +2. A major contribution to momentum diffusion coefficients appears from n = +2. However, the <span class="hlt">banded</span> structures in D alpha and Dpp coefficients appear only in the profile of diffusion coefficients for n = +2. The contribution of other resonances to diffusion coefficients is found to be, in general, quite small or even negligible. For LBC and upper <span class="hlt">band</span> chorus waves, the <span class="hlt">banded</span> structures appear only in Landau resonance. The Dpp diffusion coefficient for ECH waves is one to two orders smaller than D alpha coefficients. For chorus waves, Dpp coefficients are about an order of magnitude smaller than D alpha coefficients for the case n does not = 0. In case of Landau resonance, the values of Dpp coefficient are generally larger than the values of D alpha coefficients particularly at lower <span class="hlt">energies</span>. As an aid to the interpretation of results, we have also determined the resonant frequencies. For ECH waves, resonant frequencies have been estimated for wave normal angle 89 deg and harmonic resonances n = +1, +2, and +3, whereas for whistler mode waves, the frequencies have been calculated for angle</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19830053420&hterms=palestine&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dpalestine','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19830053420&hterms=palestine&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dpalestine"><span>The pulse profile of the Crab pulsar in the <span class="hlt">energy</span> range 45 <span class="hlt">keV</span>-1.2 MeV</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Wilson, R. B.; Fishman, G. J.</p> <p>1983-01-01</p> <p>The Crab Nebula pulsar (PSR 0531+21) is the best studied and most intense of the nontransient X-ray pulsars. However, since its spectrum drops rapidly with <span class="hlt">energy</span>, a well-resolved pulse profile has not previously been obtained above 200 <span class="hlt">keV</span>. In the hard X-ray and low-<span class="hlt">energy</span> gamma-ray region, an accurate pulse profile can be obtained with a balloon-borne detector of sufficient area during a single transit of the source. A new measurement of the pulse profile of PSR 0531+21 in the <span class="hlt">energy</span> range above 45 <span class="hlt">keV</span> obtained with a large-area scintillation detector array is reported. The detector array was flown on a balloon launched from Palestine, Texas on 1980 October 6, reaching a float altitude 4.5 g/sq cm at 0230 UTC October 7. The primary objective of the experiment was to detect and study weak gamma-ray bursts.</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('http://adsabs.harvard.edu/abs/2018PhRvA..97c2706Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PhRvA..97c2706Y"><span>Collision cross sections of N2 by H+ impact at <span class="hlt">keV</span> <span class="hlt">energies</span> within time-dependent density-functional theory</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yu, W.; Gao, C.-Z.; Zhang, Y.; Zhang, F. S.; Hutton, R.; Zou, Y.; Wei, B.</p> <p>2018-03-01</p> <p>We calculate electron capture and ionization cross sections of N2 impacted by the H+ projectile at <span class="hlt">keV</span> <span class="hlt">energies</span>. To this end, we employ the time-dependent density-functional theory coupled nonadiabatically to molecular dynamics. To avoid the explicit treatment of the complex density matrix in the calculation of cross sections, we propose an approximate method based on the assumption of constant ionization rate over the period of the projectile passing the absorbing boundary. Our results agree reasonably well with experimental data and semi-empirical results within the measurement uncertainties in the considered <span class="hlt">energy</span> range. The discrepancies are mainly attributed to the inadequate description of exchange-correlation functional and the crude approximation for constant ionization rate. Although the present approach does not predict the experiments quantitatively for collision <span class="hlt">energies</span> below 10 <span class="hlt">keV</span>, it is still helpful to calculate total cross sections of ion-molecule collisions within a certain <span class="hlt">energy</span> range.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/11573801','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/11573801"><span>The response of a thermoluminescent dosimeter to low <span class="hlt">energy</span> protons in the range 30-100 <span class="hlt">keV</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Chu, T C; Lin, S Y; Hsu, C C; Li, J P</p> <p>2001-11-01</p> <p>This study demonstrates the thermoluminescence (TL) response of CaF2:Tm (commercial name TLD-300) to 30-100 <span class="hlt">keV</span> protons which were generated by means of a Cockcroft-Walton accelerator. The phenomenon in which the total thermoluminescent output from CaF2:Tm (TLD-300) decreases with proton <span class="hlt">energy</span> from 30 to 100 <span class="hlt">keV</span> (with increase of LET) can be interpreted by the track structure theory (TST). The analysis of the glow peaks: P2 (131 degrees C), P3 (153.5 degrees C) and P6 (259 degrees C), of TLD-300 show the oscillatory decreasing phenomenon as a function of incident proton <span class="hlt">energy</span>, which can be interpreted with the TST and the oscillatory emission of electrons in a thermoluminescent dosimeter (TLD) that is caused by resonant or quasi-resonant charge transfer in ion-atom interactions in this TLD-300.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1988PhRvC..38.1083G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1988PhRvC..38.1083G"><span>Systematic features in the structure of doubly odd nuclei around A~=80 mass region: <span class="hlt">Band</span> structure in 76Rb</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>García Bermúdez, G.; Baktash, C.; Lister, C. J.; Cardona, M. A.</p> <p>1988-08-01</p> <p>Multiple-particle γ-ray coincidence techniques have been used to establish the high spin structure of 76Rb. Two ΔI=1 <span class="hlt">bands</span> were found built on the Iπ=1- ground state and on the Iπ=(4+) isomeric state at 316.8 <span class="hlt">keV</span> <span class="hlt">energy</span>. Systematic of positive parity <span class="hlt">bands</span> seen in the Br-Kr-Rb isotones with N=39 and 41 is discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20170003516&hterms=1089&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3D%2526%25231089','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20170003516&hterms=1089&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3D%2526%25231089"><span><span class="hlt">Banded</span> Structures in Electron Pitch Angle Diffusion Coefficients from Resonant Wave-Particle Interactions</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Tripathi, A. K.; Singhal, R. P.; Khazanov, G. V.; Avanov, L. A.</p> <p>2016-01-01</p> <p>Electron pitch angle (D(sub (alpha alpha))) and momentum (D(sub pp)) diffusion coefficients have been calculated due to resonant interactions with electrostatic electron cyclotron harmonic (ECH) and whistler mode chorus waves. Calculations have been performed at two spatial locations L=4.6 and 6.8 for electron <span class="hlt">energies</span> less than or equal to 10 <span class="hlt">keV</span>. Landau (n=0) resonance and cyclotron harmonic resonances n= +/- 1, +/-2, ... +/-5 have been included in the calculations. It is found that diffusion coefficient versus pitch angle (alpha) profiles show large dips and oscillations or <span class="hlt">banded</span> structures. The structures are more pronounced for ECH and lower <span class="hlt">band</span> chorus (LBC) and particularly at location 4.6. Calculations of diffusion coefficients have also been performed for individual resonances. It is noticed that the main contribution of ECH waves in pitch angle diffusion coefficient is due to resonances n=+1 and n=+2. A major contribution to momentum diffusion coefficients appears from n=+2. However, the <span class="hlt">banded</span> structures in D(sub alpha alpha) and D(sub pp) coefficients appear only in the profile of diffusion coefficients for n=+2. The contribution of other resonances to diffusion coefficients is found to be, in general, quite small or even negligible. For LBC and upper <span class="hlt">band</span> chorus waves, the <span class="hlt">banded</span> structures appear only in Landau resonance. The D(sub pp) diffusion coefficient for ECH waves is one to two orders smaller than D(sub alpha alpha) coefficients. For chorus waves, D(sub pp) coefficients are about an order of magnitude smaller than D(sub alpha alpha) coefficients for the case n does not equal 0. In case of Landau resonance, the values of D(sub pp) coefficient are generally larger than the values of D(sub alpha alpha) coefficients particularly at lower <span class="hlt">energies</span>. As an aid to the interpretation of results, we have also determined the resonant frequencies. For ECH waves, resonant frequencies have been estimated for wave normal angle 89 deg and harmonic resonances</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28918330','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28918330"><span>Thick target total bremsstrahlung spectra of lead compounds in the photon <span class="hlt">energy</span> region 1-10<span class="hlt">keV</span> by 90Sr beta particles.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Sharma, Suhansar Jit; Singh, Tajinder; Singh, Doordarshi; Singh, Amrit; Dhaliwal, A S</p> <p>2017-12-01</p> <p>Total bremsstrahlung spectral photon distribution generated in thick targets of lead compounds Pb(CH 3 COO) 2 ·3H 2 O, Pb(NO 3 ) 2 and PbCl 2 by 90 Sr beta particles has been investigated theoretically and experimentally in the photon <span class="hlt">energy</span> region 1-10<span class="hlt">keV</span>. The experimental results are compared with the theoretical models describing ordinary bremsstrahlung and the theoretical model which includes polarization bremsstrahlung into ordinary bremsstrahlung, in stripped approximation. It is observed that the experimental results show better agreement with the model which describes bremsstrahlung in stripped approximation in the <span class="hlt">energy</span> range 3-10<span class="hlt">keV</span>. However, the results show positive deviation in the photon <span class="hlt">energy</span> region of 1-3<span class="hlt">keV</span>. Further, it has been found that there is a continuous decrease of polarization bremsstrahlung contribution into ordinary bremsstrahlung in the formation of total bremsstrahlung spectra with increase in photon <span class="hlt">energy</span>. The suppression of polarization bremsstrahlung has been observed due to the presence of large fraction of low Z elements in the compounds. The results clearly indicate that polarization bremsstrahlung plays an important role in the formation of total bremsstrahlung spectra in compounds in the studied <span class="hlt">energy</span> region. Copyright © 2017 Elsevier Ltd. All rights reserved.</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> </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/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('http://adsabs.harvard.edu/abs/2011RaPC...80.1316V','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011RaPC...80.1316V"><span>Electron momentum density and <span class="hlt">band</span> structure calculations of α- and β-GeTe</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Vadkhiya, Laxman; Arora, Gunjan; Rathor, Ashish; Ahuja, B. L.</p> <p>2011-12-01</p> <p>We have measured isotropic experimental Compton profile of α-GeTe by employing high <span class="hlt">energy</span> (662 <span class="hlt">keV</span>) γ-radiation from a 137Cs isotope. To compare our experiment, we have also computed <span class="hlt">energy</span> <span class="hlt">bands</span>, density of states, electron momentum densities and Compton profiles of α- and β-phases of GeTe using the linear combination of atomic orbitals method. The electron momentum density is found to play a major role in understanding the topology of <span class="hlt">bands</span> in the vicinity of the Fermi level. It is seen that the density functional theory (DFT) with generalised gradient approximation is relatively in better agreement with the experiment than the local density approximation and hybrid Hartree-Fock/DFT.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JInst..12P4009S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JInst..12P4009S"><span>A new spectroscopic imager for X-rays from 0.5 <span class="hlt">keV</span> to 150 <span class="hlt">keV</span> combining a pnCCD and a columnar CsI(Tl) scintillator</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Schlosser, D. M.; Hartmann, R.; Kalok, D.; Bechteler, A.; Abboud, A.; Shokr, M.; Çonka, T.; Pietsch, U.; Strüder, L.</p> <p>2017-04-01</p> <p>By combining a low noise fully depleted pnCCD detector with a columnar CsI(Tl) scintillator an <span class="hlt">energy</span> dispersive spatial resolving detector can be realized with a high quantum efficiency in the range from below 0.5 <span class="hlt">keV</span> to above 150 <span class="hlt">keV</span>. The used scintillator system increases the pulse height of gamma-rays converted in the CsI(Tl), due to focusing properties of the columnar scintillator structure by reducing the event size in indirect detection mode (conversion in the scintillator). In case of direct detection (conversion in the silicon of the pnCCD) the relative <span class="hlt">energy</span> resolution is 0.7% at 122 <span class="hlt">keV</span> (FWHM = 850 eV) and the spatial resolution is less than 75 μm. In case of indirect detection the relative <span class="hlt">energy</span> resolution, integrated over all event sizes is about 9% at 122 <span class="hlt">keV</span> with an expected spatial precision of below 75 μm.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22053785-novel-flat-response-ray-detector-photon-energy-range-kev','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22053785-novel-flat-response-ray-detector-photon-energy-range-kev"><span>A novel flat-response x-ray detector in the photon <span class="hlt">energy</span> range of 0.1-4 <span class="hlt">keV</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>Li Zhichao; Guo Liang; Research Center of Laser Fusion, China Academy of Engineering Physics, Mianyang, Sichuan 621900</p> <p>2010-07-15</p> <p>A novel flat-response x-ray detector has been developed for the measurement of radiation flux from a hohlraum. In order to obtain a flat response in the photon <span class="hlt">energy</span> range of 0.1-4 <span class="hlt">keV</span>, it is found that both the cathode and the filter of the detector can be made of gold. A further improvement on the compound filter can then largely relax the requirement of the calibration x-ray beam. The calibration of the detector, which is carried out on Beijing Synchrotron Radiation Facility at Institute of High <span class="hlt">Energy</span> Physics, shows that the detector has a desired flat response in the photonmore » <span class="hlt">energy</span> range of 0.1-4 <span class="hlt">keV</span>, with a response flatness smaller than 13%. The detector has been successfully applied in the hohlraum experiment on Shenguang-III prototype laser facility. The radiation temperatures inferred from the detector agree well with those from the diagnostic instrument Dante installed at the same azimuth angle from the hohlraum axis, demonstrating the feasibility of the detector.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/20687719','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/20687719"><span>A novel flat-response x-ray detector in the photon <span class="hlt">energy</span> range of 0.1-4 <span class="hlt">keV</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Li, Zhichao; Jiang, Xiaohua; Liu, Shenye; Huang, Tianxuan; Zheng, Jian; Yang, Jiamin; Li, Sanwei; Guo, Liang; Zhao, Xuefeng; Du, Huabin; Song, Tianming; Yi, Rongqing; Liu, Yonggang; Jiang, Shaoen; Ding, Yongkun</p> <p>2010-07-01</p> <p>A novel flat-response x-ray detector has been developed for the measurement of radiation flux from a hohlraum. In order to obtain a flat response in the photon <span class="hlt">energy</span> range of 0.1-4 <span class="hlt">keV</span>, it is found that both the cathode and the filter of the detector can be made of gold. A further improvement on the compound filter can then largely relax the requirement of the calibration x-ray beam. The calibration of the detector, which is carried out on Beijing Synchrotron Radiation Facility at Institute of High <span class="hlt">Energy</span> Physics, shows that the detector has a desired flat response in the photon <span class="hlt">energy</span> range of 0.1-4 <span class="hlt">keV</span>, with a response flatness smaller than 13%. The detector has been successfully applied in the hohlraum experiment on Shenguang-III prototype laser facility. The radiation temperatures inferred from the detector agree well with those from the diagnostic instrument Dante installed at the same azimuth angle from the hohlraum axis, demonstrating the feasibility of the detector.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4512723','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4512723"><span>Possible Contrast Media Reduction with Low <span class="hlt">keV</span> Monoenergetic Images in the Detection of Focal Liver Lesions: A Dual-<span class="hlt">Energy</span> CT Animal Study</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Chung, Yong Eun; You, Je Sung; Lee, Hye-Jeong; Lim, Joon Seok; Lee, Hye Sun; Baek, Song-Ee; Kim, Myeong-Jin</p> <p>2015-01-01</p> <p>Objective To investigate the feasibility of dual-<span class="hlt">energy</span> CT for contrast media (CM) reduction in the diagnosis of hypervascular and hypovascular focal liver lesions (FLL). Subjects and Methods The Institutional Animal Care and Use Committee approved this study. VX2 tumors were implanted in two different segments of the liver in 13 rabbits. After 2 weeks, two phase contrast enhanced CT scans including the arterial phase (AP) and portal-venous phase (PVP) were performed three times with 24-hour intervals with three different concentrations of iodine, 300 (I300), 150 (I150) and 75 mg I/mL (I75). The mean HU and standard deviation (SD) were measured in the liver, the hypervascular portion of the VX2 tumor which represented hypervascular tumors, and the central necrotic area of the VX2 tumor which represented hypovascular tumors in 140kVp images with I300 as a reference standard and in monoenergetic images (between 40<span class="hlt">keV</span> and 140<span class="hlt">keV</span>) with I150 and I75. The contrast-to-noise ratio (CNR) for FLLs and the ratio of the CNRs (CNRratio) between monoenergetic image sets with I150 and I75, and the reference standard were calculated. Results For hypervascular lesions, the CNRratio was not statistically different from 1.0 between 40<span class="hlt">keV</span> and 70<span class="hlt">keV</span> images with I150, whereas the CNRratio was significantly lower than 1.0 in all <span class="hlt">keV</span> images with I75. For hypovascular lesions, the CNRratio was similar to or higher than 1.0 between 40<span class="hlt">keV</span> and 80<span class="hlt">keV</span> with I150 and between 40<span class="hlt">keV</span> and 70<span class="hlt">keV</span> with I75. Conclusions With dual-<span class="hlt">energy</span> CT, the total amount of CM might be halved in the diagnosis of hypervascular FLLs and reduced to one-fourth in the diagnosis of hypovascular FLLs, while still preserving CNRs. PMID:26203652</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 additional 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 <span class="hlt">keV</span> 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('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 <span class="hlt">keV</span> 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> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2001JPhB...34..613K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2001JPhB...34..613K"><span>L x-ray production cross sections in Th and U at 17.8, 25.8 and 46.9 <span class="hlt">keV</span> photon <span class="hlt">energies</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kumar, Ajay; Puri, Sanjiv; Shahi, J. S.; Garg, M. L.; Mehta, D.; Singh, Nirmal</p> <p>2001-02-01</p> <p>The L x-ray production (XRP) differential cross sections in Th and U have been measured at the 17.8 <span class="hlt">keV</span> incident photon <span class="hlt">energy</span> (E_L3<Einc<E_L2, ELi is the Li subshell ionization threshold) in an angular range 90°-160°, and at the 25.8 and 46.9 <span class="hlt">keV</span> incident photon <span class="hlt">energies</span> (E_L1<Einc<EK) at an angle of 130°. The measurements were performed using the <span class="hlt">energy</span> dispersive x-ray fluorescence set-up in secondary excitation mode. In contrast to the findings of an earlier experiment (Sharma and Allawadhi 1999 J. Phys. B: At. Mol. Opt. Phys. 32 2343), the present measurements rule out the possibility of a strong angular dependence of differential cross sections for various L3 subshell x-rays following selective photoionization of the L3 subshell. Integral L XRP cross sections at the 17.8, 25.8 and 46.9 <span class="hlt">keV</span> photon <span class="hlt">energies</span>, deduced assuming isotropic emission of the L x-rays, are found to be in good agreement with those evaluated using the most reliable theoretical values of Li (i = 1,2,3) subshell photoionization cross sections, fluorescence yields, x-ray emission rates and Coster-Kronig transition probabilities.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2007SPIE.6705E..12I','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2007SPIE.6705E..12I"><span>High-<span class="hlt">energy</span>-resolution monochromator for nuclear resonant scattering of synchrotron radiation by Te-125 at 35.49 <span class="hlt">keV</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Imai, Yasuhiko; Yoda, Yoshitaka; Kitao, Shinji; Masuda, Ryo; Higashitaniguchi, Satoshi; Inaba, Chika; Seto, Makoto</p> <p>2007-09-01</p> <p>We have developed a high-resolution monochromator (HRM) for the measurement of nuclear resonant scattering (NRS) of synchrotron radiation by Te-125 at 35.49 <span class="hlt">keV</span> using the backscattering of sapphire (9 1 -10 68). HRMs for nuclei with excitation <span class="hlt">energies</span> less than 30 <span class="hlt">keV</span> have been successfully developed using high angle diffractions by silicon crystals. Nearly perfect silicon crystal, however, is not suitable for high efficient HRMs at higher <span class="hlt">energy</span> regions because the symmetry of the crystal structure is high and the Debye-temperature is low. Therefore, we used high quality synthetic sapphire crystal, which has low symmetry of crystal structure and high Debye-temperature. The temperature of the crystal was precisely controlled around 218 K to diffract synchrotron radiation with a Bragg angle of π/2 - 0.52 mrad. <span class="hlt">Energy</span> was tuned by changing the crystal temperature under the condition of constant diffraction angle. <span class="hlt">Energy</span> resolution was measured by detecting nuclear forward scattering by Te-125 in enriched TeO II. The relative <span class="hlt">energy</span> resolution of 2.1×10 -7 is achieved, that is 7.5 meV in <span class="hlt">energy</span> bandwidth. This HRM opens studies on element-specific dynamics and electronic state of substances containing Te-125.</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('https://www.ncbi.nlm.nih.gov/pubmed/18697550','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/18697550"><span>Dose distribution of a 125 <span class="hlt">keV</span> mean <span class="hlt">energy</span> microplanar x-ray beam for basic studies on microbeam radiotherapy.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Ohno, Yumiko; Torikoshi, Masami; Suzuki, Masao; Umetani, Keiji; Imai, Yasuhiko; Uesugi, Kentaro; Yagi, Naoto</p> <p>2008-07-01</p> <p>A multislit collimator was designed and fabricated for basic studies on microbeam radiation therapy (MRT) with an x-ray <span class="hlt">energy</span> of about 100 <span class="hlt">keV</span>. It consists of 30 slits that are 25 microm high, 30 mm wide, and 5 mm thick in the beam direction. The slits were made of 25 microm-thick polyimide sheets that were separated by 175 microm-thick tungsten sheets. The authors measured the dose distribution of a single microbeam with a mean <span class="hlt">energy</span> of 125 <span class="hlt">keV</span> by a scanning slit method using a phosphor coupled to a charge coupled device camera and found that the ratios of the dose at the center of a microbeam to that at midpositions to adjacent slits were 1050 and 760 for each side of the microbeam. This dose distribution was well reproduced by the Monte Carlo simulation code PHITS.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/21255389-search-anomalous-scattering-ev-kev','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/21255389-search-anomalous-scattering-ev-kev"><span>Search For Anomalous n-p Scattering At 60 eV-140 <span class="hlt">keV</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>Moreh, R.; Block, R. C.; Danon, Y.</p> <p>2009-01-28</p> <p>A search for an anomalous n-p scattering from a polyethylene sample (CH{sub 2}) at 8 final <span class="hlt">energies</span> between 64 eV and 2.5 <span class="hlt">keV</span> was carried out. The scattering intensities were compared to that from a graphite (C) sample. The results were found to confirm our previous n-p results on H{sub 2}O at a final <span class="hlt">energy</span> of 24.3 <span class="hlt">keV</span> where no n-p scattering anomaly was observed. The present results refute all proposed models attempting to explain the occurrence of any n-p scattering anomaly at <span class="hlt">keV</span> neutron <span class="hlt">energies</span>.</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 addition, 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 <span class="hlt">keV</span>), 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/22016338-survival-depth-organics-ices-under-low-energy-electron-radiation-kev','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22016338-survival-depth-organics-ices-under-low-energy-electron-radiation-kev"><span>SURVIVAL DEPTH OF ORGANICS IN ICES UNDER LOW-<span class="hlt">ENERGY</span> ELECTRON RADIATION ({<=}2 <span class="hlt">keV</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>Barnett, Irene Li; Lignell, Antti; Gudipati, Murthy S., E-mail: gudipati@jpl.nasa.gov</p> <p>2012-03-01</p> <p>Icy surfaces in our solar system are continually modified and sputtered with electrons, ions, and photons from solar wind, cosmic rays, and local magnetospheres in the cases of Jovian and Saturnian satellites. In addition to their prevalence, electrons specifically are expected to be a principal radiolytic agent on these satellites. Among energetic particles (electrons and ions), electrons penetrate by far the deepest into the ice and could cause damage to organic material of possible prebiotic and even biological importance. To determine if organic matter could survive and be detected through remote sensing or in situ explorations on these surfaces, suchmore » as water ice-rich Europa, it is important to obtain accurate data quantifying electron-induced chemistry and damage depths of organics at varying incident electron <span class="hlt">energies</span>. Experiments reported here address the quantification issue at lower electron <span class="hlt">energies</span> (100 eV-2 <span class="hlt">keV</span>) through rigorous laboratory data analysis obtained using a novel methodology. A polycyclic aromatic hydrocarbon molecule, pyrene, embedded in amorphous water ice films of controlled thicknesses served as an organic probe. UV-VIS spectroscopic measurements enabled quantitative assessment of organic matter survival depths in water ice. Eight ices of various thicknesses were studied to determine damage depths more accurately. The electron damage depths were found to be linear, approximately 110 nm keV{sup -1}, in the tested range which is noticeably higher than predictions by Monte Carlo simulations by up to 100%. We conclude that computational simulations underestimate electron damage depths in the <span class="hlt">energy</span> region {<=}2 <span class="hlt">keV</span>. If this trend holds at higher electron <span class="hlt">energies</span> as well, present models utilizing radiation-induced organic chemistry in icy solar system bodies need to be revisited. For interstellar ices of a few micron thicknesses, we conclude that low-<span class="hlt">energy</span> electrons generated through photoionization processes in the interstellar</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2004JAP....95.4117B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2004JAP....95.4117B"><span>Charge dynamics of MgO single crystals subjected to <span class="hlt">KeV</span> electron irradiation</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Boughariou, A.; Blaise, G.; Braga, D.; Kallel, A.</p> <p>2004-04-01</p> <p>A scanning electron microscope has been equipped to study the fundamental aspects of charge trapping in insulating materials, by measuring the secondary electron emission (SEE) yield σ with a high precision (a few percent), as a function of <span class="hlt">energy</span>, electron current density, and dose. The intrinsic secondary electron emission yield σ0 of uncharged MgO single crystals annealed at 1000 °C, 2 h, has been studied at four <span class="hlt">energies</span> 1.1, 5, 15, and 30 <span class="hlt">keV</span> on three different crystal orientations (100), (110), and (111). At low <span class="hlt">energies</span> (1.1 and 5 <span class="hlt">keV</span>) σ0 depends on the crystalline orientation wheras at high <span class="hlt">energies</span> (30 <span class="hlt">keV</span>) no differentiation occurs. It is shown that the value of the second crossover <span class="hlt">energy</span> E2, for which the intrinsic SEE yield σ0=1, is extremely delicate to measure with precision. It is about 15 keV±500 eV for the (100) orientation, 13.5 keV±500 eV for the (110), and 18.5 keV±500 eV for the (111) one. At low current density J⩽105 pA/cm2, the variation of σ with the injected dose makes possible the observation of a self-regulated regime characterized by a steady value of the SEE yield σst=1. At low <span class="hlt">energies</span> 1.1 and 5 <span class="hlt">keV</span>, there is no current density effects in MgO, but at high <span class="hlt">energies</span> ≈30 <span class="hlt">keV</span>, apparent current density effects come from a bad collect of secondary electrons, due to very high negative surface potential. At 30 <span class="hlt">keV</span> <span class="hlt">energy</span>, an intense erratic electron exoemission was observed on the MgO (110) orientation annealed at 1500 °C. This phenomenon is the result of a disruptive process similar to flashover, which takes place at the surface of the material.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018RaPC..146...26K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018RaPC..146...26K"><span>An analytical X-ray CdTe detector response matrix for incomplete charge collection correction for photon <span class="hlt">energies</span> up to 300 <span class="hlt">keV</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kurková, Dana; Judas, Libor</p> <p>2018-05-01</p> <p>Gamma and X-ray <span class="hlt">energy</span> spectra measured with semiconductor detectors suffer from various distortions, one of them being so-called "tailing" caused by an incomplete charge collection. Using the Hecht equation, a response matrix of size 321 × 321 was constructed which was used to correct the effect of incomplete charge collection. The correction matrix was constructed analytically for an arbitrary <span class="hlt">energy</span> bin and the size of the <span class="hlt">energy</span> bin thus defines the width of the spectral window. The correction matrix can be applied separately from other possible spectral corrections or it can be incorporated into an already existing response matrix of the detector. The correction was tested and its adjustable parameters were optimized on the line spectra of 57Co measured with a cadmium telluride (CdTe) detector in a spectral range from 0 up to 160 <span class="hlt">keV</span>. The best results were obtained when the values of the free path of holes were spread over a range from 0.4 to 1.0 cm and weighted by a Gauss function. The model with the optimized parameter values was then used to correct the line spectra of 152Eu in a spectral range from 0 up to 530 <span class="hlt">keV</span>. An improvement in the <span class="hlt">energy</span> resolution at full width at half maximum from 2.40 % ± 0.28 % to 0.96 % ± 0.28 % was achieved at 344.27 <span class="hlt">keV</span>. Spectra of "narrow spectrum series" beams, N120, N150, N200, N250 and N300, generated with tube voltages of 120 kV, 150 kV, 200 kV, 250 kV and 300 kV respectively, and measured with the CdTe detector, were corrected in the spectral range from 0 to 160 <span class="hlt">keV</span> (N120 and N150) and from 0 to 530 <span class="hlt">keV</span> (N200, N250, N300). All the measured spectra correspond both qualitatively and quantitatively to the available reference data after the correction. To obtain better correspondence between N150, N200, N250 and N300 spectra and the reference data, lower values of the free paths of holes (range from 0.16 to 0.65 cm) were used for X-ray spectra correction, which suggests <span class="hlt">energy</span> dependence of the phenomenon.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26971414','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26971414"><span>Carotid dual-<span class="hlt">energy</span> CT angiography: Evaluation of low <span class="hlt">keV</span> calculated monoenergetic datasets by means of a frequency-split approach for noise reduction at low <span class="hlt">keV</span> levels.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Riffel, Philipp; Haubenreisser, Holger; Meyer, Mathias; Sudarski, Sonja; Morelli, John N; Schmidt, Bernhard; Schoenberg, Stefan O; Henzler, Thomas</p> <p>2016-04-01</p> <p>Calculated monoenergetic ultra-low <span class="hlt">keV</span> datasets did not lead to improved contrast-to-noise ratio (CNR) due to the dramatic increase in image noise. The aim of the present study was to evaluate the objective image quality of ultra-low <span class="hlt">keV</span> monoenergetic images (MEIs) calculated from carotid DECT angiography data with a new monoenergetic imaging algorithm using a frequency-split technique. 20 patients (12 male; mean age 53±17 years) were retrospectively analyzed. MEIs from 40 to 120 <span class="hlt">keV</span> were reconstructed using the monoenergetic split frequency approach (MFSA). Additionally MEIs were reconstructed for 40 and 50 <span class="hlt">keV</span> using a conventional monoenergetic (CM) software application. Signal intensity, noise, signal-to-noise ratio (SNR) and CNR were assessed in the basilar, common, internal carotid arteries. Ultra-low <span class="hlt">keV</span> MEIs at 40 <span class="hlt">keV</span> and 50 <span class="hlt">keV</span> demonstrated highest vessel attenuation, significantly greater than those of the polyenergetic images (PEI) (all p-values <0.05). The highest SNR level and CNR level was found at 40 <span class="hlt">keV</span> and 50 <span class="hlt">keV</span> (all p-values <0.05). MEIs with MFSA showed significantly lower noise levels than those processed with CM (all p-values <0.05) and no significant differences in vessel attenuation (p>0.05). Thus MEIs with MFSA showed significantly higher SNR and CNR compared to MEIs with CM. Combining the lower spatial frequency stack for contrast at low <span class="hlt">keV</span> levels with the high spatial frequency stack for noise at high <span class="hlt">keV</span> levels (frequency-split technique) leads to improved image quality of ultra-low <span class="hlt">keV</span> monoenergetic DECT datasets when compared to previous monoenergetic reconstruction techniques without the frequency-split technique. Copyright © 2016 Elsevier Ireland Ltd. All rights reserved.</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('https://www.osti.gov/biblio/21120854-dose-distribution-kev-mean-energy-microplanar-ray-beam-basic-studies-microbeam-radiotherapy','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/21120854-dose-distribution-kev-mean-energy-microplanar-ray-beam-basic-studies-microbeam-radiotherapy"><span>Dose distribution of a 125 <span class="hlt">keV</span> mean <span class="hlt">energy</span> microplanar x-ray beam for basic studies on microbeam radiotherapy</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>Ohno, Yumiko; Torikoshi, Masami; Suzuki, Masao</p> <p></p> <p>A multislit collimator was designed and fabricated for basic studies on microbeam radiation therapy (MRT) with an x-ray <span class="hlt">energy</span> of about 100 <span class="hlt">keV</span>. It consists of 30 slits that are 25 {mu}m high, 30 mm wide, and 5 mm thick in the beam direction. The slits were made of 25 {mu}m-thick polyimide sheets that were separated by 175 {mu}m-thick tungsten sheets. The authors measured the dose distribution of a single microbeam with a mean <span class="hlt">energy</span> of 125 <span class="hlt">keV</span> by a scanning slit method using a phosphor coupled to a charge coupled device camera and found that the ratios of themore » dose at the center of a microbeam to that at midpositions to adjacent slits were 1050 and 760 for each side of the microbeam. This dose distribution was well reproduced by the Monte Carlo simulation code PHITS.« 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_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('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('http://adsabs.harvard.edu/abs/2017EPJWC.14601014G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EPJWC.14601014G"><span>Neutron capture cross sections of 69Ga and 71Ga at 25 <span class="hlt">keV</span> and Epeak = 90 <span class="hlt">keV</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Göbel, Kathrin; Beinrucker, Clemens; Erbacher, Philipp; Fiebiger, Stefan; Fonseca, Micaela; Heftrich, Michael; Heftrich, Tanja; Käppeler, Franz; Krása, Antonin; Lederer-Woods, Claudia; Plag, Ralf; Plompen, Arjan; Reifarth, René; Schmidt, Stefan; Sonnabend, Kerstin; Weigand, Mario</p> <p>2017-09-01</p> <p>We measured the neutron capture cross sections of 69Ga and 71Ga for a quasi-stellar spectrum at kBT = 25 <span class="hlt">keV</span> and a spectrum with a peak <span class="hlt">energy</span> at 90 <span class="hlt">keV</span> by the activation technique at the Joint Research Centre (JRC) in Geel, Belgium. Protons were provided by an electrostatic Van de Graaff accelerator to produce neutrons via the reaction 7Li(p,n). The produced activity was measured via the γ emission of the product nuclei by high-purity germanium detectors. We present preliminary results.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20140000880','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20140000880"><span>Heliospheric Neutral Atom Spectra Between 0.01 and 6 <span class="hlt">keV</span> fom IBEX</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Fuselier, S. A.; Allegrini, F.; Bzowski, M.; Funsten, H. O.; Ghielmetti, A. G.; Gloeckler, G.; Heirtzler, D.; Janzen, P.; Kubiak, M.; Kucharek, H.; <a style="text-decoration: none; " href="javascript:void(0); " onClick="displayelement('author_20140000880'); toggleEditAbsImage('author_20140000880_show'); toggleEditAbsImage('author_20140000880_hide'); "> <img style="display:inline; width:12px; height:12px; " src="images/arrow-up.gif" width="12" height="12" border="0" alt="hide" id="author_20140000880_show"> <img style="width:12px; height:12px; display:none; " src="images/arrow-down.gif" width="12" height="12" border="0" alt="hide" id="author_20140000880_hide"></p> <p>2012-01-01</p> <p>Since 2008 December, the Interstellar Boundary Explorer (IBEX) has been making detailed observations of neutrals from the boundaries of the heliosphere using two neutral atom cameras with overlapping <span class="hlt">energy</span> ranges. The unexpected, yet defining feature discovered by IBEX is a Ribbon that extends over the <span class="hlt">energy</span> range from about 0.2 to 6 <span class="hlt">keV</span>. This Ribbon is superposed on a more uniform, globally distributed heliospheric neutral population. With some important exceptions, the focus of early IBEX studies has been on neutral atoms with <span class="hlt">energies</span> greater than approx. 0.5 <span class="hlt">keV</span>. With nearly three years of science observations, enough low-<span class="hlt">energy</span> neutral atom measurements have been accumulated to extend IBEX observations to <span class="hlt">energies</span> less than approx. 0.5 <span class="hlt">keV</span>. Using the <span class="hlt">energy</span> overlap of the sensors to identify and remove backgrounds, <span class="hlt">energy</span> spectra over the entire IBEX <span class="hlt">energy</span> range are produced. However, contributions by interstellar neutrals to the <span class="hlt">energy</span> spectrum below 0.2 <span class="hlt">keV</span> may not be completely removed. Compared with spectra at higher <span class="hlt">energies</span>, neutral atom spectra at lower <span class="hlt">energies</span> do not vary much from location to location in the sky, including in the direction of the IBEX Ribbon. Neutral fluxes are used to show that low <span class="hlt">energy</span> ions contribute approximately the same thermal pressure as higher <span class="hlt">energy</span> ions in the heliosheath. However, contributions to the dynamic pressure are very high unless there is, for example, turbulence in the heliosheath with fluctuations of the order of 50-100 km/s.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017PhRvC..96c4321G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017PhRvC..96c4321G"><span>Outstanding problems in the <span class="hlt">band</span> structures of 152Sm</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.; Hamilton, J. H.</p> <p>2017-09-01</p> <p>The recent data on B (E 2 ) values, deduced from the multi-Coulex excitation of the low spin states in the decay of 152Sm, and other experimental findings in the last two decades are compared with the predictions from the microscopic dynamic pairing plus quadrupole model of Kumar and Baranger. The 1292.8 <span class="hlt">keV</span> 2+ state is assigned to the 03 + <span class="hlt">band</span>, and the K =2 assignment of the 1769 <span class="hlt">keV</span> 2+ state is confirmed. The anomaly of the shape coexistence of the assumed spherical β <span class="hlt">band</span> versus the deformed ground <span class="hlt">band</span> is resolved. The values from the critical point symmetry X(5) support the collective character of the β <span class="hlt">band</span>. The problem with the two-term interacting boson model Hamiltonian in predicting β and γ <span class="hlt">bands</span> in 152Sm leads to interesting consequences. The collective features of the second excited Kπ=03 + <span class="hlt">band</span> are preferred over the "pairing isomer" view. Also the multiphonon nature of the higher lying Kπ=22 +β γ <span class="hlt">band</span> and Kπ=4+ <span class="hlt">band</span> are illustrated vis-à-vis the new data and the nuclear structure theory.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JInst..12C1016M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JInst..12C1016M"><span>Testing and Comparison of Imaging Detectors for Electrons in the <span class="hlt">Energy</span> Range 10-20 <span class="hlt">keV</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Matheson, J.; Moldovan, G.; Kirkland, A.; Allinson, N.; Abrahams, J. P.</p> <p>2017-11-01</p> <p>Interest in direct detectors for low-<span class="hlt">energy</span> electrons has increased markedly in recent years. Detection of electrons in the <span class="hlt">energy</span> range up to low tens of <span class="hlt">keV</span> is important in techniques such as photoelectron emission microscopy (PEEM) and electron backscatter diffraction (EBSD) on scanning electron microscopes (SEMs). The PEEM technique is used both in the laboratory and on synchrotron light sources worldwide. The ubiquity of SEMs means that there is a very large market for EBSD detectors for materials studies. Currently, the most widely used detectors in these applications are based on indirect detection of incident electrons. Examples include scintillators or microchannel plates (MCPs), coupled to CCD cameras. Such approaches result in blurring in scintillators/phosphors, distortions in optical systems, and inefficiencies due the limited active area of MCPs. In principle, these difficulties can be overcome using direct detection in a semiconductor device. Growing out of a feasibility study into the use of a direct detector for use on an XPEEM, we have built at Rutherford Appleton Laboratory a system to illuminate detectors with an electron beam of <span class="hlt">energy</span> up to 20 <span class="hlt">keV</span> . We describe this system in detail. It has been used to measure the performance of a custom back-thinned monolithic active pixel sensor (MAPS), a detector based on the Medipix2 chip, and a commercial detector based on MCPs. We present a selection of the results from these measurements and compare and contrast different detector types.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19930031011&hterms=soft+power&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3Dsoft%2Bpower','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19930031011&hterms=soft+power&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3Dsoft%2Bpower"><span>The soft X-ray diffuse background observed with the HEAO 1 low-<span class="hlt">energy</span> detectors</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Garmire, G. P.; Nousek, J. A.; Apparao, K. M. V.; Burrows, D. N.; Fink, R. L.; Kraft, R. P.</p> <p>1992-01-01</p> <p>Results of a study of the diffuse soft-X-ray background as observed by the low-<span class="hlt">energy</span> detectors of the A-2 experiment aboard the HEAO 1 satellite are reported. The observed sky intensities are presented as maps of the diffuse X-ray background sky in several <span class="hlt">energy</span> <span class="hlt">bands</span> covering the <span class="hlt">energy</span> range 0.15-2.8 <span class="hlt">keV</span>. It is found that the soft X-ray diffuse background (SXDB) between 1.5 and 2.8 <span class="hlt">keV</span>, assuming a power law form with photon number index 1.4, has a normalization constant of 10.5 +/- 1.0 photons/sq cm s sr <span class="hlt">keV</span>. Below 1.5 <span class="hlt">keV</span> the spectrum of the SXDB exceeds the extrapolation of this power law. The low-<span class="hlt">energy</span> excess for the NEP can be fitted with emission from a two-temperature equilibrium plasma model with the temperatures given by log I1 = 6.16 and log T2 = 6.33. It is found that this model is able to account for the spectrum below 1 <span class="hlt">keV</span>, but fails to yield the observed Galactic latitude variation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23192280','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23192280"><span>Measurement of the mass <span class="hlt">energy</span>-absorption coefficient of air for x-rays in the range from 3 to 60 <span class="hlt">keV</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Buhr, H; Büermann, L; Gerlach, M; Krumrey, M; Rabus, H</p> <p>2012-12-21</p> <p>For the first time the absolute photon mass <span class="hlt">energy</span>-absorption coefficient of air in the <span class="hlt">energy</span> range of 10 to 60 <span class="hlt">keV</span> has been measured with relative standard uncertainties below 1%, considerably smaller than those of up to 2% assumed for calculated data. For monochromatized synchrotron radiation from the electron storage ring BESSY II both the radiant power and the fraction of power deposited in dry air were measured using a cryogenic electrical substitution radiometer and a free air ionization chamber, respectively. The measured absorption coefficients were compared with state-of-the art calculations and showed an average deviation of 2% from calculations by Seltzer. However, they agree within 1% with data calculated earlier by Hubbell. In the course of this work, an improvement of the data analysis of a previous experimental determination of the mass <span class="hlt">energy</span>-absorption coefficient of air in the range of 3 to 10 <span class="hlt">keV</span> was found to be possible and corrected values of this preceding study are given.</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('https://www.osti.gov/servlets/purl/1415390','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1415390"><span>Emittance Effects on Gain in $W$ -<span class="hlt">Band</span> TWTs</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>Carlsten, Bruce Eric; Nichols, Kimberley E.; Shchegolkov, Dmitry Yu.</p> <p></p> <p>We consider the main effects of beam emittance on W-<span class="hlt">band</span> traveling-wave tube (TWT) performance and gain. Specifically, we consider a representative dielectric TWT structure with ~5 dB/cm of gain driven by a 5-A, 20-<span class="hlt">keV</span>, sheet electron beam that is focused by a wiggler magnetic field. The normalized beam transverse emittance must be about 1 μm or lower to ensure that both the transport is stable and the gain is not degraded by the effective <span class="hlt">energy</span> spread arising from the emittance. This emittance limit scales roughly inversely with frequency.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1415390-emittance-effects-gain-band-twts','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1415390-emittance-effects-gain-band-twts"><span>Emittance Effects on Gain in $W$ -<span class="hlt">Band</span> TWTs</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Carlsten, Bruce Eric; Nichols, Kimberley E.; Shchegolkov, Dmitry Yu.; ...</p> <p>2016-10-20</p> <p>We consider the main effects of beam emittance on W-<span class="hlt">band</span> traveling-wave tube (TWT) performance and gain. Specifically, we consider a representative dielectric TWT structure with ~5 dB/cm of gain driven by a 5-A, 20-<span class="hlt">keV</span>, sheet electron beam that is focused by a wiggler magnetic field. The normalized beam transverse emittance must be about 1 μm or lower to ensure that both the transport is stable and the gain is not degraded by the effective <span class="hlt">energy</span> spread arising from the emittance. This emittance limit scales roughly inversely with frequency.</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 Additional, Hard Power-Law Component from 10 <span class="hlt">keV</span> 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 » <span class="hlt">keV</span> 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 additional 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.osti.gov/servlets/purl/944334','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/944334"><span>Absolute Calibration of Image Plate for electrons at <span class="hlt">energy</span> between 100 <span class="hlt">keV</span> and 4 MeV</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>Chen, H; Back, N L; Eder, D C</p> <p>2007-12-10</p> <p>The authors measured the absolute response of image plate (Fuji BAS SR2040) for electrons at <span class="hlt">energies</span> between 100 <span class="hlt">keV</span> to 4 MeV using an electron spectrometer. The electron source was produced from a short pulse laser irradiated on the solid density targets. This paper presents the calibration results of image plate Photon Stimulated Luminescence PSL per electrons at this <span class="hlt">energy</span> range. The Monte Carlo radiation transport code MCNPX results are also presented for three representative incident angles onto the image plates and corresponding electron <span class="hlt">energies</span> depositions at these angles. These provide a complete set of tools that allows extraction ofmore » the absolute calibration to other spectrometer setting at this electron <span class="hlt">energy</span> range.« less</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('https://www.osti.gov/pages/biblio/1351137-electron-elevator-excitations-across-band-gap-via-dynamical-gap-state','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1351137-electron-elevator-excitations-across-band-gap-via-dynamical-gap-state"><span>Electron elevator: Excitations across the <span class="hlt">band</span> gap via a dynamical gap state</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Lim, Anthony; Foulkes, W. M. C.; Horsfield, A. P.; ...</p> <p>2016-01-27</p> <p>We use time-dependent density functional theory to study self-irradiated Si. We calculate the electronic stopping power of Si in Si by evaluating the <span class="hlt">energy</span> transferred to the electrons per unit path length by an ion of kinetic <span class="hlt">energy</span> from 1 eV to 100 <span class="hlt">keV</span> moving through the host. Electronic stopping is found to be significant below the threshold velocity normally identified with transitions across the <span class="hlt">band</span> gap. A structured crossover at low velocity exists in place of a hard threshold. Lastly, an analysis of the time dependence of the transition rates using coupled linear rate equations enables one of themore » excitation mechanisms to be clearly identified: a defect state induced in the gap by the moving ion acts like an elevator and carries electrons across the <span class="hlt">band</span> gap.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018ApJ...853...89Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018ApJ...853...89Y"><span>The Strongest Acceleration of >40 <span class="hlt">keV</span> Electrons by ICME-driven Shocks at 1 au</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yang, Liu; Wang, Linghua; Li, Gang; Wimmer-Schweingruber, Robert F.; He, Jiansen; Tu, Chuanyi; Tian, Hui; Bale, Stuart D.</p> <p>2018-01-01</p> <p>We present two case studies of the in-situ electron acceleration during the 2000 February 11 shock and the 2004 July 22 shock, with the strongest electron flux enhancement at 40 <span class="hlt">keV</span> across the shock, among all the quasi-perpendicular and quasi-parallel ICME-driven shocks observed by the WIND 3DP instrument from 1995 through 2014 at 1 au. We find that for this quasi-perpendicular (quasi-parallel) shock on 2000 February 11 (2004 July 22), the shocked electron differential fluxes at ∼0.4–50 <span class="hlt">keV</span> in the downstream generally fit well to a double-power-law spectrum, J ∼ E ‑β , with an index of β ∼ 3.15 (4.0) at <span class="hlt">energies</span> below a break at ∼3 <span class="hlt">keV</span> (∼1 <span class="hlt">keV</span>) and β ∼ 2.65 (2.6) at <span class="hlt">energies</span> above. For both shock events, the downstream electron spectral indices appear to be similar for all pitch angles, which are significantly larger than the index prediction by diffusive shock acceleration. In addition, the downstream electron pitch-angle distributions show the anisotropic beams in the anti-sunward-traveling direction, while the ratio of the downstream over ambient fluxes appears to peak near 90° pitch angles, at all <span class="hlt">energies</span> of ∼0.4–50 <span class="hlt">keV</span>. These results suggest that in both shocks, shock drift acceleration likely plays an important role in accelerating electrons in situ at 1 au. Such ICME-driven shocks could contribute to the formation of solar wind halo electrons at <span class="hlt">energies</span> ≲2 <span class="hlt">keV</span>, as well as the production of solar wind superhalo electrons at <span class="hlt">energies</span> ≳2 <span class="hlt">keV</span> in interplanetary space.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JGRA..122.9282C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JGRA..122.9282C"><span>Dominance of high-<span class="hlt">energy</span> (>150 <span class="hlt">keV</span>) heavy ion intensities in Earth's middle to outer magnetosphere</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Cohen, Ian J.; Mitchell, Donald G.; Kistler, Lynn M.; Mauk, Barry H.; Anderson, Brian J.; Westlake, Joseph H.; Ohtani, Shinichi; Hamilton, Douglas C.; Turner, Drew L.; Blake, J. Bernard; Fennell, Joseph F.; Jaynes, Allison N.; Leonard, Trevor W.; Gerrard, Andrew J.; Lanzerotti, Louis J.; Allen, Robert C.; Burch, James L.</p> <p>2017-09-01</p> <p>Previous observations have driven the prevailing assumption in the field that energetic ions measured by an instrument using a bare solid state detector (SSD) are predominantly protons. However, new near-equatorial energetic particle observations obtained between 7 and 12 RE during Phase 1 of the Magnetospheric Multiscale mission challenge the validity of this assumption. In particular, measurements by the Energetic Ion Spectrometer (EIS) instruments have revealed that the intensities of heavy ion species (specifically oxygen and helium) dominate those of protons at <span class="hlt">energies</span> ≳150-220 <span class="hlt">keV</span> in the middle to outer (>7 RE) magnetosphere. Given that relative composition measurements can drift as sensors degrade in gain, quality cross-calibration agreement between EIS observations and those from the SSD-based Fly's Eye Energetic Particle Spectrometer (FEEPS) sensors provides critical support to the veracity of the measurement. Similar observations from the Radiation Belt Storm Probes Ion Composition Experiment (RBSPICE) instruments aboard the Van Allen Probes spacecraft extend the ion composition measurements into the middle magnetosphere and reveal a strongly proton-dominated environment at L≲6 but decreasing proton intensities at L≳6. It is concluded that the intensity dominance of the heavy ions at higher <span class="hlt">energies</span> (>150 <span class="hlt">keV</span>) arises from the existence of significant populations of multiply-charged heavy ions, presumably of solar wind origin.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014RaPC..102...49B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014RaPC..102...49B"><span>Linear attenuation coefficients of tissues from 1 <span class="hlt">keV</span> to 150 <span class="hlt">keV</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Böke, Aysun</p> <p>2014-09-01</p> <p>The linear attenuation coefficients and three interaction processes have been computed for liver, kidney, muscle, fat and for a range of x-ray <span class="hlt">energies</span> from 1 <span class="hlt">keV</span> to 150 <span class="hlt">keV</span>. Molecular photoelectric absorption cross sections were calculated from atomic cross section data. Total coherent (Rayleigh) and incoherent (Compton) scattering cross sections were obtained by numerical integration over combinations of F2m(x) with the Thomson formula and Sm(x) with the Klein-Nishina formula, respectively. For the coherent (Rayleigh) scattering cross section calculations, molecular form factors were obtained from recent experimental data in the literature for values of x<1 Å-1 and from the relativistic modified atomic form factors for values of x≥1 Å-1. With the inclusion of molecular interference effects in the coherent (Rayleigh) scattering, more accurate knowledge of the scatter from these tissues will be provided. The number of elements involved in tissue composition is 5 for liver, 47 for kidney, 44 for muscle and 3 for fat. The results are compared with previously published experimental and theoretical linear attenuation coefficients. In general, good agreement is obtained. The molecular form factors and scattering functions and cross sections are incorporated into a Monte Carlo program. The <span class="hlt">energy</span> distributions of x-ray photons scattered from tissues have been simulated and the results are presented.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFMSM52A..09M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFMSM52A..09M"><span>Diffusive transport of several hundred <span class="hlt">keV</span> electrons in the Earth's slot region</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ma, Q.; Li, W.; Thorne, R. M.; Bortnik, J.</p> <p>2017-12-01</p> <p>We investigate the gradual diffusion of energetic electrons from the inner edge of the outer radiation belt into the slot region. The Van Allen Probes observed slow inward diffusion and decay of 200-600 <span class="hlt">keV</span> electrons following the intense geomagnetic storm that occurred on 17 March 2013. During the 10-day non-disturbed period following the storm, the peak of electron fluxes gradually moved from L 2.7 to L 2.4, and the flux levels decreased by a factor of 2-4 depending on the electron <span class="hlt">energy</span>. We simulated the radial intrusion and decay of electrons using a 3-dimentional diffusion code, which reproduced the <span class="hlt">energy</span>-dependent transport of electrons from 100 <span class="hlt">keV</span> to 1 MeV in the slot region. At <span class="hlt">energies</span> of 100-200 <span class="hlt">keV</span>, the electrons experience fast transport across the slot region due to the dominance of radial diffusion; at <span class="hlt">energies</span> of 200-600 <span class="hlt">keV</span>, the electrons gradually diffuse and decay in the slot region due to the comparable radial diffusion rate and pitch angle scattering rate by plasmaspheric hiss; at <span class="hlt">energies</span> of E > 700 <span class="hlt">keV</span>, the electrons stopped diffusing near the inner edge of outer radiation belt due to the dominant pitch angle scattering loss. In addition to plasmaspheric hiss, magnetosonic waves and VLF waves can cause the loss of high pitch angle electrons, relaxing the sharp `top-hat' shaped pitch angle distributions created by plasmaspheric hiss. Our simulation indicates the importance of radial diffusion and pitch angle scattering in forming the diffusive intrusion of energetic electrons across the slot region.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2006RScI...77jF308M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2006RScI...77jF308M"><span>Absolute calibration of Kodak Biomax-MS film response to x rays in the 1.5- to 8-<span class="hlt">keV</span> <span class="hlt">energy</span> range</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Marshall, F. J.; Knauer, J. P.; Anderson, D.; Schmitt, B. L.</p> <p>2006-10-01</p> <p>The absolute response of Kodak Biomax-MS film to x rays in the range from 1.5- to 8-<span class="hlt">keV</span> has been measured using a laboratory electron-beam generated x-ray source. The measurements were taken at specific line <span class="hlt">energies</span> by using Bragg diffraction to produce monochromatic beams of x rays. Multiple exposures were taken on Biomax MS film up to levels exceeding optical densities of 2 as measured by a microdensitometer. The absolute beam intensity for each exposure was measured with a Si (Li) detector. Additional response measurements were taken with Kodak direct exposure film (DEF) so as to compare the results of this technique to previously published calibrations. The Biomax-MS results have been fitted to a semiempirical mathematical model (Knauer et al., these proceedings). Users of the model can infer absolute fluences from observed exposure levels at either interpolated or extrapolated <span class="hlt">energies</span>. To summarize the results: Biomax MS has comparable sensitivity to DEF film below 3<span class="hlt">keV</span> but has reduced sensitivity above 3<span class="hlt">keV</span> (˜50%). The lower exposure results from thinner emulsion layers, designed for use with phosphor screens. The ease with which Biomax-MS can be used in place of DEF (same format film, same developing process, and comparable sensitivity) makes it a good replacement.</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> </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://ntrs.nasa.gov/search.jsp?R=19930055634&hterms=Krieger&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3DKrieger','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19930055634&hterms=Krieger&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3DKrieger"><span>Multilayer diffraction at 104 <span class="hlt">keV</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Krieger, Allen S.; Blake, Richard L.; Siddons, D. P.</p> <p>1993-01-01</p> <p>We have measured the diffraction peak of a W:Si synthetic multilayer reflector at 104 <span class="hlt">keV</span> using the High <span class="hlt">Energy</span> Bonse-Hart Camera at the X-17B hard X-ray wiggler beam line of the National Synchrotron Light Source at Brookhaven National Laboratory. The characteristics of the diffraction peak are described and compared to theory.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017yCat..17970091M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017yCat..17970091M"><span>VizieR Online Data Catalog: X-Ray source properties for NGC 2207/IC 2163 (Mineo+, 2014)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mineo, S.; Rappaport, S.; Levine, A.; Pooley, D.; Steinhorn, B.; Homan, J.</p> <p>2017-08-01</p> <p>We analyzed four Chandra ACIS-S observations of the galaxy pair NGC 2207/IC 2163. The data reduction was done following the standard CIAO threads (CIAO version 4.6, CALDB version 4.5.9) for soft (0.5-2 <span class="hlt">keV</span>), hard (2-8 <span class="hlt">keV</span>), and broad (0.5-8.0 <span class="hlt">keV</span>) <span class="hlt">energy</span> <span class="hlt">bands</span>. All Chandra data sets were reprocessed using chandra_repro, a script that automates the recommended data-processing steps presented in the CIAO analysis threads. Using the script fluximage, we computed a monochromatic exposure map for the mean photon <span class="hlt">energy</span> of each <span class="hlt">band</span>: 1.25 <span class="hlt">keV</span>, 5.0 <span class="hlt">keV</span>, and 4.25 <span class="hlt">keV</span> for the soft, hard, and broad <span class="hlt">band</span>, respectively. fluximage outputs both the instrument map for the center of each <span class="hlt">energy</span> <span class="hlt">band</span> using the tool mkinstmap and the exposure maps in sky coordinates for each <span class="hlt">energy</span> <span class="hlt">band</span> using mkexpmap. (5 data files).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016RaPC..123..122R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016RaPC..123..122R"><span>Measurement of Lα and Lβ1,3,4 fluorescence cross sections of La, Ce, Pr and Nd induced by photons of <span class="hlt">energies</span> between 7.01 <span class="hlt">keV</span> and 8.75 <span class="hlt">keV</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Reyes-Herrera, J.; Miranda, J.</p> <p>2016-06-01</p> <p>This study presents measurement results of x-ray production cross sections of Lα and Lβ1,3,4 emitted by four lanthanoid elements (La, Ce, Pr and Nd), after irradiation with Kα and Kβ X rays of the elements Co, Ni, Cu, and Zn (covering <span class="hlt">energies</span> between 7.01 <span class="hlt">keV</span> and 8.75 <span class="hlt">keV</span>). Primary x-rays were induced in turn by the irradiation of thick targets of these elements with a beam of x-rays produced by a tube with an Rh anode, operating at 50 kV and 850 μA. The experimental results are compared with theoretical cross sections predicted using known tabulations of photoelectric cross sections. Dirac-Hartree-Slater (DHS) atomic parameters were used for these calculations. An acceptable match between experiment and both sets of tabulated data is found.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2007APS..MARH25010P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2007APS..MARH25010P"><span>Biological Response of Cancer and Normal Cells on Irradiation from Electrons with <span class="hlt">Energies</span> up to 200 <span class="hlt">keV</span>.</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Prilepskiy, Yuriy</p> <p>2007-03-01</p> <p>This paper presents continuation data of the series of experiments with the electron gun of the CEBAF machine at Jefferson Lab (Newport News, VA), which is capable of delivering electrons with <span class="hlt">energies</span> up to 200 <span class="hlt">keV</span>. This 1.5 GHz beam permits to generate cellular damage within minutes. We have performed irradiation of cancer and normal cells with different electron <span class="hlt">energies</span> and currents to investigate cell biological responses. The biological response is measured through proteomics analysis before and after irradiation. The living cells are encased in special air containers allowing proper positioning in vacuum where the electrons are present. The containers receive the irradiation from the mono energetic electrons with <span class="hlt">energy</span> up to 120 <span class="hlt">keV</span>, resulting in an irradiation from both electrons and a small number of photons from the original beam passing through the thin container window. This window allows approximately half of the beam to come through. The study will permit to address the physical processes involved in the RBE and LET at a level that supersedes current data listed in the literature. We will discuss the experimental setup and the second stage of data collected with the new more developed system. This research is part of a global program to provide detailed information for the understanding of radiation based cancer treatments.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017PhRvM...1e5201C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017PhRvM...1e5201C"><span>Ionoluminescence properties of polystyrene-hosted fluorophore films induced by helium ions of <span class="hlt">energy</span> 50-350 <span class="hlt">keV</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Chakraborty, Subha; Huang, Mengbing</p> <p>2017-10-01</p> <p>We report on measurements and analysis of ionoluminescence properties of pure polystyrene films and polystyrene films doped with four types of fluorophores in low kinetic <span class="hlt">energies</span> (50-350 <span class="hlt">keV</span>) of ion irradiation. We have developed a theoretical model to understand the experimentally observed ionoluminescence behaviors in terms of scintillation yield from individual ion tracks, photophysical <span class="hlt">energy</span> transfer mechanisms, and irradiation-induced defects. A comparison of the model and experimental results suggests that singlet up-conversion resulting from triplet-triplet annihilation processes may be responsible for enhanced singlet emission of the fluorophores at high ion beam flux densities. <span class="hlt">Energy</span> transfer from the polystyrene matrix to the fluorophore molecules has been identified as an effective pathway to increasing the fluorescence efficiency in the doped scintillator films.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017ATel10733....1P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017ATel10733....1P"><span>Swift/BAT sees MAXI J1535-571 declining in 15-50 <span class="hlt">keV</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Palmer, D. M.; Krimm, H. A.; Swift/BAT Team</p> <p>2017-09-01</p> <p>BAT monitoring of MAXI J1535-571 shows that, in the 15-50 <span class="hlt">keV</span> <span class="hlt">band</span>, the source reached a peak around 2017 Sept 9 and has since begun to decline. At peak, the BAT count rate was 0.43 +/- 0.015 ct s-1 cm-2, or approximately twice the mean flux of the Crab.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JGRA..12210235M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JGRA..12210235M"><span>Diffusive Transport of Several Hundred <span class="hlt">keV</span> Electrons in the Earth's Slot Region</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ma, Q.; Li, W.; Thorne, R. M.; Bortnik, J.; Reeves, G. D.; Spence, H. E.; Turner, D. L.; Blake, J. B.; Fennell, J. F.; Claudepierre, S. G.; Kletzing, C. A.; Kurth, W. S.; Hospodarsky, G. B.; Baker, D. N.</p> <p>2017-10-01</p> <p>We investigate the gradual diffusion of energetic electrons from the inner edge of the outer radiation belt into the slot region. The Van Allen Probes observed slow inward diffusion and decay of 200-600 <span class="hlt">keV</span> electrons following the intense geomagnetic storm that occurred on 17 March 2013. During the 10 day nondisturbed period following the storm, the peak of electron fluxes gradually moved from L 2.7 to L 2.4, and the flux levels decreased by a factor of 2-4 depending on the electron <span class="hlt">energy</span>. We simulated the radial intrusion and decay of electrons using a three-dimensional diffusion code, which reproduced the <span class="hlt">energy</span>-dependent transport of electrons from 100 <span class="hlt">keV</span> to 1 MeV in the slot region. At <span class="hlt">energies</span> of 100-200 <span class="hlt">keV</span>, the electrons experience fast transport across the slot region due to the dominance of radial diffusion; at <span class="hlt">energies</span> of 200-600 <span class="hlt">keV</span>, the electrons gradually diffuse and decay in the slot region due to the comparable rate of radial diffusion and pitch angle scattering by plasmaspheric hiss; at <span class="hlt">energies</span> of E > 700 <span class="hlt">keV</span>, the electrons stopped diffusing near the inner edge of outer radiation belt due to the dominant pitch angle scattering loss. In addition to plasmaspheric hiss, magnetosonic waves and VLF transmitters can cause the loss of high pitch angle electrons, relaxing the sharp "top-hat" shaped pitch angle distributions created by plasmaspheric hiss. Our simulation indicates the importance of balance between radial diffusion and loss through pitch angle scattering in forming the diffusive intrusion of energetic electrons across the slot region.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018AIPC.1966b0035K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AIPC.1966b0035K"><span>Structural and optical modification in 4H-SiC following 30 <span class="hlt">keV</span> silver ion irradiation</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kaushik, Priya Darshni; Aziz, Anver; Siddiqui, Azher M.; Lakshmi, G. B. V. S.; Syväjärvi, Mikael; Yakimova, Rositsa; Yazdi, G. Reza</p> <p>2018-05-01</p> <p>The market of high power, high frequency and high temperature based electronic devices is captured by SiC due to its superior properties like high thermal conductivity and high sublimation temperature and also due to the limitation of silicon based electronics in this area. There is a need to investigate effect of ion irradiation on SiC due to its application in outer space as outer space is surrounded both by low and high <span class="hlt">energy</span> ion irradiations. In this work, effect of low <span class="hlt">energy</span> ion irradiation on structural and optical property of 4H-SiC is investigated. ATR-FTIR is used to study structural modification and UV-Visible spectroscopy is used to study optical modifications in 4H-SiC following 30 <span class="hlt">keV</span> Ag ion irradiation. FTIR showed decrease in bond density of SiC along the ion path (track) due to the creation of point defects. UV-Visible absorption spectra showed decrease in optical <span class="hlt">band</span> gap from 3.26 eV to 2.9 eV. The study showed degradation of SiC crystallity and change in optical <span class="hlt">band</span> gap following low <span class="hlt">energy</span> ion irradiation and should be addressed while fabricationg devices based on SiC for outer space application. Additionally, this study provides a platform for introducing structural and optical modification in 4H-SiC using ion beam technology in a controlled manner.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1345934-unraveling-excitation-mechanisms-highly-oblique-lower-band-chorus-waves','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1345934-unraveling-excitation-mechanisms-highly-oblique-lower-band-chorus-waves"><span>Unraveling the excitation mechanisms of highly oblique lower-<span class="hlt">band</span> chorus waves</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Li, Wen; Mourenas, D.; Artemyev, A. V.; ...</p> <p>2016-08-17</p> <p>Excitation mechanisms of highly oblique, quasi-electrostatic lower <span class="hlt">band</span> chorus waves are investigated using Van Allen Probes observations near the equator of the Earth's magnetosphere. Linear growth rates are evaluated based on in situ, measured electron velocity distributions and plasma conditions and compared with simultaneously observed wave frequency spectra and wave normal angles. Accordingly, two distinct excitation mechanisms of highly oblique lower <span class="hlt">band</span> chorus have been clearly identified for the first time. The first mechanism relies on cyclotron resonance with electrons possessing both a realistic temperature anisotropy at <span class="hlt">keV</span> <span class="hlt">energies</span> and a plateau at 100–500 eV in the parallel velocity distribution.more » The second mechanism corresponds to Landau resonance with a 100–500 eV beam. In both cases, a small low-<span class="hlt">energy</span> beam-like component is necessary for suppressing an otherwise dominating Landau damping. In conclusion, our new findings suggest that small variations in the electron distribution could have important impacts on energetic electron dynamics.« less</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 <span class="hlt">keV</span>, 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/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('https://www.osti.gov/biblio/20861338-absolute-calibration-kodak-biomax-ms-film-response-rays-kev-energy-range','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/20861338-absolute-calibration-kodak-biomax-ms-film-response-rays-kev-energy-range"><span>Absolute calibration of Kodak Biomax-MS film response to x rays in the 1.5- to 8-<span class="hlt">keV</span> <span class="hlt">energy</span> range</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>Marshall, F. J.; Knauer, J. P.; Anderson, D.</p> <p>2006-10-15</p> <p>The absolute response of Kodak Biomax-MS film to x rays in the range from 1.5- to 8-<span class="hlt">keV</span> has been measured using a laboratory electron-beam generated x-ray source. The measurements were taken at specific line <span class="hlt">energies</span> by using Bragg diffraction to produce monochromatic beams of x rays. Multiple exposures were taken on Biomax MS film up to levels exceeding optical densities of 2 as measured by a microdensitometer. The absolute beam intensity for each exposure was measured with a Si(Li) detector. Additional response measurements were taken with Kodak direct exposure film (DEF) so as to compare the results of this techniquemore » to previously published calibrations. The Biomax-MS results have been fitted to a semiempirical mathematical model (Knauer et al., these proceedings). Users of the model can infer absolute fluences from observed exposure levels at either interpolated or extrapolated <span class="hlt">energies</span>. To summarize the results: Biomax MS has comparable sensitivity to DEF film below 3 <span class="hlt">keV</span> but has reduced sensitivity above 3 <span class="hlt">keV</span> ({approx}50%). The lower exposure results from thinner emulsion layers, designed for use with phosphor screens. The ease with which Biomax-MS can be used in place of DEF (same format film, same developing process, and comparable sensitivity) makes it a good replacement.« less</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 additional 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/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('http://adsabs.harvard.edu/abs/2005SPIE.5650...64Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2005SPIE.5650...64Y"><span>Optimization of single <span class="hlt">keV</span> ion implantation for the construction of single P-donor devices</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yang, Changyi; Jamieson, David N.; Hopf, Toby; Andresen, Soren E.; Hearne, Sean M.; Hudson, Fay E.; Pakes, Christopher I.; Mitic, Mladen; Gauja, Eric; Tamanyan, Grigori; Dzurak, Andrew S.; Prawer, Steven; Clark, Robert G.</p> <p>2005-02-01</p> <p>We report recent progress in single <span class="hlt">keV</span> ion implantation and online detection for the controlled implantation of single donors in silicon. When integrated with silicon nanofabrication technology this forms the "top down" strategy for the construction of prototype solid state quantum computer devices based on phosphorus donors in silicon. We have developed a method of single ion implantation and online registration that employs detector electrodes adjacent to the area into which the donors are to be implanted. The implantation sites are positioned with nanometer accuracy using an electron beam lithography patterned PMMA mask. Control of the implantation depth of 20 nm is achieved by tuning the phosphorus ion <span class="hlt">energy</span> to 14 <span class="hlt">keV</span>. The counting of single ion implantation in each site is achieved by the detection of e-/h+ pairs produced by the implanted phosphorus ion in the substrate. The system is calibrated by use of Mn K-line x-rays (5.9 and 6.4 <span class="hlt">keV</span>) and we find the ionization <span class="hlt">energy</span> of the 14 <span class="hlt">keV</span> phosphorus ions in silicon to be about 3.5-4.0 <span class="hlt">keV</span> for implants through a 5 nm SiO2 surface layer. This paper describes the development of an improved PIN detector structure that provides more reliable performance of the earlier MOS structure. With the new structure, the <span class="hlt">energy</span> noise threshold has been minimized to 1 <span class="hlt">keV</span> or less. Unambiguous detection/counting of single <span class="hlt">keV</span> ion implantation events were achieved with a confidence level greater than 98% with a reliable and reproducible fabrication process.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20140017688','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20140017688"><span>Decline of the 2-10 <span class="hlt">keV</span> Emission from Eta Carinae</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Liburd, Jamar; Corcoran, Michael F.; Hamaguchi, Kenji; Gull, Theodore R.; Madura, Thomas; Teodoro, Mairan; Moffat, Anthony; Richardson, Noel; Russell, Chris; Pollock, Andrew; <a style="text-decoration: none; " href="javascript:void(0); " onClick="displayelement('author_20140017688'); toggleEditAbsImage('author_20140017688_show'); toggleEditAbsImage('author_20140017688_hide'); "> <img style="display:inline; width:12px; height:12px; " src="images/arrow-up.gif" width="12" height="12" border="0" alt="hide" id="author_20140017688_show"> <img style="width:12px; height:12px; display:none; " src="images/arrow-down.gif" width="12" height="12" border="0" alt="hide" id="author_20140017688_hide"></p> <p>2014-01-01</p> <p>Analysis of Eta Car's X-ray spectrum in the 2-10 <span class="hlt">keV</span> <span class="hlt">band</span> using processed data from the X-ray Telescope on Swift reveals a peak flux on July 16, 2014 of 0.046 photons s(exp -1) cm(exp -2) (3.37+/-0.15×10(exp -10) ergs s(exp -1) cm(exp -2). This flux is similar to the previous maximum flux seen by the XRT, 3.53+/-0.13×10(exp -10) ergs s(exp -1) cm(exp -2) (0.049 photons s(exp -1) cm(exp -2), ATEL #6298). Since this peak on July 16, the most recent Swift XRT quicklook data show a drop in flux. On July 20, 2014 the XRT flux as seen in the quicklook data was 0.011 photons s(exp -1) cm(exp -2) (8.3+/-0.5×10(exp -11) ergs s(exp -1) cm(exp -2)). This most likely indicates that the 2-10 <span class="hlt">keV</span> flux is in its declining phase as Eta Car approaches its deep X-ray minimum stage (Hamaguchi et al., 2014, ApJ, 784, 125) associated with periastron passage of the 2024-day binary orbit. The column density derived from analysis of the July 20 XRT quicklook data is 7.2×10(exp 22) cm(exp -2). This is consistent with the column density seen near the same orbital phase in 2003 (7.7×10(exp 22) cm(exp -2), Hamaguchi et al., 2007, ApJ, 663, 522). Eta Car's deep X-ray minimum phase is expected to begin on July 30, 2014. Weekly Swift/XRT observations of Eta Car in the 2-10 <span class="hlt">keV</span> <span class="hlt">band</span> are planned throughout the X-ray minimum.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22590996-measurement-mass-attenuation-coefficient-from-kev-kev-elemental-materials-al-cu-pb','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22590996-measurement-mass-attenuation-coefficient-from-kev-kev-elemental-materials-al-cu-pb"><span>Measurement of the mass attenuation coefficient from 81 <span class="hlt">keV</span> to 1333 <span class="hlt">keV</span> for elemental materials Al, Cu and Pb</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>Gjorgieva, Slavica, E-mail: slavicagjorgieva89@gmail.com; Institute of Physics, Faculty of Natural Sciences and Mathematics, Ss Cyril and Methodius University, POB 162, 1000 Skopje; Barandovski, Lambe, E-mail: lambe@pmf.ukim.mk</p> <p></p> <p>The mass attenuation coefficients (μ/ρ) for 3 high purity elemental materials Al, Cu and Pb were measured in the γ-ray <span class="hlt">energy</span> range from 81 <span class="hlt">keV</span> up to 1333 <span class="hlt">keV</span> using {sup 22}Na, {sup 60}Co {sup 133}Ba and {sup 133}Cs as sources of gamma radiation. Well shielded detector (NaI (Tl) semiconductor detector) was used to measure the intensity of the transmitted beam. The measurements were made under condition of good geometry, assuring that any photon absorbed or deflected appreciably does not reach the detector. The measured values are compared with the theoretical ones obtained by Seltzer (1993).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1440490-diffusive-transport-several-hundred-kev-electrons-earth-slot-region','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1440490-diffusive-transport-several-hundred-kev-electrons-earth-slot-region"><span>Diffusive Transport of Several Hundred <span class="hlt">keV</span> Electrons in the Earth's Slot Region</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Ma, Q.; Li, W.; Thorne, R. M.; ...</p> <p>2017-09-29</p> <p>Here, we investigate the gradual diffusion of energetic electrons from the inner edge of the outer radiation belt into the slot region. The Van Allen Probes observed slow inward diffusion and decay of ~200–600 <span class="hlt">keV</span> electrons following the intense geomagnetic storm that occurred on 17 March 2013. During the 10 day nondisturbed period following the storm, the peak of electron fluxes gradually moved from L ~ 2.7 to L ~ 2.4, and the flux levels decreased by a factor of ~2–4 depending on the electron <span class="hlt">energy</span>. We simulated the radial intrusion and decay of electrons using a three–dimensional diffusion code,more » which reproduced the energy–dependent transport of electrons from ~100 <span class="hlt">keV</span> to 1 MeV in the slot region. At <span class="hlt">energies</span> of 100–200 <span class="hlt">keV</span>, the electrons experience fast transport across the slot region due to the dominance of radial diffusion; at <span class="hlt">energies</span> of 200–600 <span class="hlt">keV</span>, the electrons gradually diffuse and decay in the slot region due to the comparable rate of radial diffusion and pitch angle scattering by plasmaspheric hiss; at <span class="hlt">energies</span> of E > 700 <span class="hlt">keV</span>, the electrons stopped diffusing near the inner edge of outer radiation belt due to the dominant pitch angle scattering loss. In addition to plasmaspheric hiss, magnetosonic waves and VLF transmitters can cause the loss of high pitch angle electrons, relaxing the sharp “top–hat” shaped pitch angle distributions created by plasmaspheric hiss. Our simulation indicates the importance of balance between radial diffusion and loss through pitch angle scattering in forming the diffusive intrusion of energetic electrons across the slot region.« less</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('https://ntrs.nasa.gov/search.jsp?R=19930009522&hterms=Day+George&qs=N%3D0%26Ntk%3DAuthor-Name%26Ntx%3Dmode%2Bmatchall%26Ntt%3DDay%252C%2BGeorge%2BS','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19930009522&hterms=Day+George&qs=N%3D0%26Ntk%3DAuthor-Name%26Ntx%3Dmode%2Bmatchall%26Ntt%3DDay%252C%2BGeorge%2BS"><span>The broad-<span class="hlt">band</span> x ray spectral variability of Mkn 841</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>George, I. M.; Nandra, K.; Fabian, A. C.; Turner, T. J.; Done, C.; Day, C. S. R.</p> <p>1992-01-01</p> <p>The results of a detailed spectral analysis of four X-ray observations of the luminous Seyfert 1.5 galaxy Mkn 841 performed using the EXOSAT and Ginga satellites over the period June 1984 to July 1990 are reported. Preliminary results from a short ROSAT PSPC observation of Mkn 841 in July 1990 are also presented. Variability is apparent in both the soft (0.1-1.0 <span class="hlt">keV</span>) and medium (1-20 <span class="hlt">keV</span>) <span class="hlt">energy</span> <span class="hlt">bands</span>. Above 1 <span class="hlt">keV</span>, the spectra are adequately modelled by a power-law with a strong emission line of equivalent width approximately 450 eV. The <span class="hlt">energy</span> of the line (approximately 6.4 <span class="hlt">keV</span>) is indicative of K-shell fluorescence from neutral iron, leading to the interpretation that the line arises via X-ray illumination of cold material surrounding the source. In addition to the flux variability, the continuum shape also changes in a dramatic fashion, with variations in the apparent photon index Delta(Gamma) approximately 0.6. The large equivalent width of the emission line clearly indicates a strongly enhanced reflection component in the source, compared to other Seyferts observed with Ginga. The spectral changes are interpreted in terms of a variable power-law continuum superimposed on a flatter reflection component. For one Ginga observation, the reflected flux appears to dominate the medium <span class="hlt">energy</span> X-ray emission, resulting in an unusually flat slope (Gamma approximately 1.0). The soft X-ray excess is found to be highly variable by a factor approximately 10. These variations are not correlated with the hard flux, but it seems likely that the soft component arises via reprocessing of the hard X-rays. We find no evidence for intrinsic absorption, with the equivalent hydrogen column density constrained to be less than or equal to few x 10(exp 20) cm(exp -2). The implications of these results for physical models for the emission regions in this and other X-ray bright Seyferts are briefly discussed.</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..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/2017ExA....44..359D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017ExA....44..359D"><span>The Cryogenic AntiCoincidence detector for ATHENA X-IFU: a scientific assessment of the observational capabilities in the hard X-ray <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>D'Andrea, M.; Lotti, S.; Macculi, C.; Piro, L.; Argan, A.; Gatti, F.</p> <p>2017-12-01</p> <p>ATHENA is a large X-ray observatory, planned to be launched by ESA in 2028 towards an L2 orbit. One of the two instruments of the payload is the X-IFU: a cryogenic spectrometer based on a large array of TES microcalorimeters, able to perform integral field spectrography in the 0.2-12 <span class="hlt">keV</span> <span class="hlt">band</span> (2.5 eV FWHM at 6 <span class="hlt">keV</span>). The X-IFU sensitivity is highly degraded by the particle background expected in the L2 orbit, which is induced by primary protons of both galactic and solar origin, and mostly by secondary electrons. To reduce the particle background level and enable the mission science goals, the instrument incorporates a Cryogenic AntiCoincidence detector (CryoAC). It is a 4 pixel TES based detector, placed < 1 mm below the main array. In this paper we report a scientific assessment of the CryoAC observational capabilities in the hard X-ray <span class="hlt">band</span> (E > 10 <span class="hlt">keV</span>). The aim of the study has been to understand if the present detector design can be improved in order to enlarge the X-IFU scientific capability on an <span class="hlt">energy</span> <span class="hlt">band</span> wider than the TES array. This is beyond the CryoAC baseline, being this instrument aimed to operate as anticoincidence particle detector and not conceived to perform X-ray observations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19950053099&hterms=Einstein&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3DEinstein','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19950053099&hterms=Einstein&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3DEinstein"><span>An Einstein survey of the 1 <span class="hlt">keV</span> soft X-ray background in the Galactic plane</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Stanford, John M.; Caillault, Jean-Pierre</p> <p>1994-01-01</p> <p>We have analyzed 56 Einstein Observatory Imaging Proportional Counter (IPC) observations within +/- 3 deg of the Galactic plane in order to determine the low-latitude soft X-ray background flux in the 0.56-1.73 <span class="hlt">keV</span> <span class="hlt">band</span>. Any detected X-ray point source which fell within our regions of study was removed from the image, enabling us to present maps of the background flux as a function of Galactic latitude along 18 meridians. These maps reveal considerable structure to the background in the Galactic plane on an angular scale of approximately 1 deg. Our results are compared with those of an earlier study of the 1 <span class="hlt">keV</span> X-ray background along l = 25 deg by Kahn & Caillault. The double-peaked structure they found is not discernible in our results, possibly because of the presence of solar backscattered flux in their data. A model which takes into account contributions to the background by extragalactic and stellar sources, the distribution of both atomic and molecular absorbing material with the Galaxy, the <span class="hlt">energy</span> dependence of the cross section for absorption of X-rays, and the <span class="hlt">energy</span> dependence of the detector has been constructed and fitted to these new data to derive constraints on the scale height, temperature, and volume emissivity of the unaccounted-for X-ray-emitting material. The results of this model along l = 25 deg are roughly similar to those of the model of Kahn & Caillault along the same meridian.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1987PhRvA..35.1430S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1987PhRvA..35.1430S"><span>Photon cross sections in Cu, Pt, and Au at 81 <span class="hlt">keV</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Seetharami Reddy, B.; Ramana Rao, P. V.; Premchand, K.; Parthasaradhi, K.</p> <p>1987-02-01</p> <p>Total photon cross sections in Cu, Pt, and Au are measured employing the doublet 79.623- and 80.999-<span class="hlt">keV</span> γ's of 133Ba. A dilution of the cross section by about 12% is observed at the average <span class="hlt">energy</span> of the doublet in Au due to K-edge falling in between these two <span class="hlt">energies</span>. Scofield's theoretical value in this case is seen to be underestimated by about 75% due to the use of different K-edge <span class="hlt">energies</span> in Au. However, an alternative but customarily followed procedure is to ignore the lower K-edge data of Scofield and extrapolate using upper-edge data which yield a value agreeing satisfactorily with the experimental value at 80.905 <span class="hlt">keV</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017RaPC..141....1G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017RaPC..141....1G"><span>Total electron scattering cross sections of some important biomolecules at 0.2-6.0 <span class="hlt">keV</span> <span class="hlt">energies</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gurung, Meera Devi; Ariyasinghe, W. M.</p> <p>2017-12-01</p> <p>The total electron scattering cross sections (TCS) of five nucleic bases (adenine, cytosine, guanine, thymine and uracil), phosphoric acid, three amino acids (glycine, lysine, and L-histidine), D-glucose, alpha-D-glucose, tetrahydropyran (THP), 3-hydroxytetrahydrofuran and furan have been determined in the <span class="hlt">energy</span> range 0.2-6.0 <span class="hlt">keV</span> using a simple model based on the effective atomic total electron scattering cross sections (EATCS). The reliability of the model is confirmed by comparing the determined TCS with the predictions of those by existing theoretical models.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017PhPl...24c3112C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017PhPl...24c3112C"><span>High-<span class="hlt">energy</span> (>70 <span class="hlt">keV</span>) x-ray conversion efficiency measurement on the ARC laser at the National Ignition Facility</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Chen, Hui; Hermann, M. R.; Kalantar, D. H.; Martinez, D. A.; Di Nicola, P.; Tommasini, R.; Landen, O. L.; Alessi, D.; Bowers, M.; Browning, D.; Brunton, G.; Budge, T.; Crane, J.; Di Nicola, J.-M.; Döppner, T.; Dixit, S.; Erbert, G.; Fishler, B.; Halpin, J.; Hamamoto, M.; Heebner, J.; Hernandez, V. J.; Hohenberger, M.; Homoelle, D.; Honig, J.; Hsing, W.; Izumi, N.; Khan, S.; LaFortune, K.; Lawson, J.; Nagel, S. R.; Negres, R. A.; Novikova, L.; Orth, C.; Pelz, L.; Prantil, M.; Rushford, M.; Shaw, M.; Sherlock, M.; Sigurdsson, R.; Wegner, P.; Widmayer, C.; Williams, G. J.; Williams, W.; Whitman, P.; Yang, S.</p> <p>2017-03-01</p> <p>The Advanced Radiographic Capability (ARC) laser system at the National Ignition Facility (NIF) is designed to ultimately provide eight beamlets with a pulse duration adjustable from 1 to 30 ps, and <span class="hlt">energies</span> up to 1.5 kJ per beamlet. Currently, four beamlets have been commissioned. In the first set of 6 commissioning target experiments, the individual beamlets were fired onto gold foil targets with <span class="hlt">energy</span> up to 1 kJ per beamlet at 20-30 ps pulse length. The x-ray <span class="hlt">energy</span> distribution and pulse duration were measured, yielding <span class="hlt">energy</span> conversion efficiencies of 4-9 × 10-4 for x-rays with <span class="hlt">energies</span> greater than 70 <span class="hlt">keV</span>. With greater than 3 J of such x-rays, ARC provides a high-precision x-ray backlighting capability for upcoming inertial confinement fusion and high-<span class="hlt">energy</span>-density physics experiments on NIF.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22545177-su-monte-carlo-investigation-radiation-interactions-gold-nanoparticles-water-mv-kev-kev-photon-beams','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22545177-su-monte-carlo-investigation-radiation-interactions-gold-nanoparticles-water-mv-kev-kev-photon-beams"><span>SU-E-T-46: A Monte Carlo Investigation of Radiation Interactions with Gold Nanoparticles in Water for 6 MV, 85 <span class="hlt">KeV</span> and 40 <span class="hlt">KeV</span> Photon 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>Flint, D B; O’Brien, D J; McFadden, C H</p> <p></p> <p>Purpose: To determine the effect of gold-nanoparticles (AuNPs) on <span class="hlt">energy</span> deposition in water for different irradiation conditions. Methods: TOPAS version B12 Monte Carlo code was used to simulate <span class="hlt">energy</span> deposition in water from monoenergetic 40 <span class="hlt">keV</span> and 85 <span class="hlt">keV</span> photon beams and a 6 MV Varian Clinac photon beam (IAEA phase space file, 10x10 cm{sup 2}, SSD 100 cm). For the 40 and 85 <span class="hlt">keV</span> beams, monoenergetic 2x2 mm{sup 2} parallel beams were used to irradiate a 30x30x10 µm {sup 3} water mini-phantom located at 1.5 cm depth in a 30x30x50 cm{sup 3} water phantom. 5000 AuNPs of 50 nmmore » diameter were randomly distributed inside the mini-phantom. <span class="hlt">Energy</span> deposition was scored in the mini-phantom with the AuNPs’ material set to gold and then water. For the 6 MV beam, we created another phase space (PHSP) file on the surface of a 2 mm diameter sphere located at 1.5 cm depth in the water phantom. The PHSP file consisted of all particles entering the sphere including backscattered particles. Simulations were then performed using the new PHSP as the source with the mini-phantom centered in a 2 mm diameter water sphere in vacuum. The g4em-livermore reference list was used with “EMRangeMin/EMRangeMax = 100 eV/7 MeV” and “SetProductionCutLowerEdge = 990 eV” to create the new PHSP, and “SetProductionCutLowerEdge = 100 eV” for the mini-phantom simulations. All other parameters were set as defaults (“finalRange = 100 µm”). Results: The addition of AuNPs resulted in an increase in the mini-phantom <span class="hlt">energy</span> deposition of (7.5 ± 8.7)%, (1.6 ± 8.2)%, and (−0.6 ± 1.1)% for 40 <span class="hlt">keV</span>, 85 <span class="hlt">keV</span> and 6 MV beams respectively. Conclusion: Enhanced <span class="hlt">energy</span> deposition was seen at low photon <span class="hlt">energies</span>, but decreased with increasing <span class="hlt">energy</span>. No enhancement was observed for the 6 MV beam. Future work is required to decrease the statistical uncertainties in the simulations. This research is partially supported from institutional funds from the Center for Radiation Oncology Research</p> </li> <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('https://www.ncbi.nlm.nih.gov/pubmed/24519325','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24519325"><span>Monte Carlo calculations of <span class="hlt">energy</span> deposition distributions of electrons below 20 <span class="hlt">keV</span> in protein.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Tan, Zhenyu; Liu, Wei</p> <p>2014-05-01</p> <p>The distributions of <span class="hlt">energy</span> depositions of electrons in semi-infinite bulk protein and the radial dose distributions of point-isotropic mono-energetic electron sources [i.e., the so-called dose point kernel (DPK)] in protein have been systematically calculated in the <span class="hlt">energy</span> range below 20 <span class="hlt">keV</span>, based on Monte Carlo methods. The ranges of electrons have been evaluated by extrapolating two calculated distributions, respectively, and the evaluated ranges of electrons are compared with the electron mean path length in protein which has been calculated by using electron inelastic cross sections described in this work in the continuous-slowing-down approximation. It has been found that for a given <span class="hlt">energy</span>, the electron mean path length is smaller than the electron range evaluated from DPK, but it is large compared to the electron range obtained from the <span class="hlt">energy</span> deposition distributions of electrons in semi-infinite bulk protein. The <span class="hlt">energy</span> dependences of the extrapolated electron ranges based on the two investigated distributions are given, respectively, in a power-law form. In addition, the DPK in protein has also been compared with that in liquid water. An evident difference between the two DPKs is observed. The calculations presented in this work may be useful in studies of radiation effects on proteins.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015ApJ...813L..22R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015ApJ...813L..22R"><span>First INTEGRAL Observations of V404 Cygni during the 2015 Outburst: Spectral Behavior in the 20-650 <span class="hlt">keV</span> <span class="hlt">Energy</span> Range</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Roques, Jean-Pierre; Jourdain, Elisabeth; Bazzano, Angela; Fiocchi, Mariateresa; Natalucci, Lorenzo; Ubertini, Pietro</p> <p>2015-11-01</p> <p>In 2015 June, the source V404 Cygni (= GS2023+38) underwent an extraordinary outburst. We present the results obtained during the first revolution dedicated to this target by the INTEGRAL mission and focus on the spectral behavior in the hard X-ray domain, using both SPI and IBIS instruments. The source exhibits extreme variability and reaches fluxes of several tens of Crab. However, the emission between 20 and 650 <span class="hlt">keV</span> can be understood in terms of two main components, varying on all the observable timescales, similar to what is observed in the persistent black hole system Cyg X-1. The low-<span class="hlt">energy</span> component (up to ˜200 <span class="hlt">keV</span>) presents a rather unusual shape, probably due to the intrinsic source variability. Nonetheless, a satisfactory description is obtained with a Comptonization model, if an unusually hot population of seed photons (kT0 ˜ 7 <span class="hlt">keV</span>) is introduced. Above this first component, a clear excess extending up to 400-600 <span class="hlt">keV</span> leads us to investigate a scenario where an additional (cutoff) power law could correspond to the contribution of the jet synchrotron emission, as proposed in Cyg X-1. A search for an annihilation feature did not provide any firm detection, with an upper limit of 2 × 10-4 ph cm-2 s-1 (2σ) for a narrow line centered at 511 <span class="hlt">keV</span>, on the averaged obtained spectrum. Based on observations with INTEGRAL, an ESA project with instruments and science data center funded by ESA member states (especially the PI countries: Denmark, France, Germany, Italy, Spain, and Switzerland), Czech Republic, and Poland with the participation of Russia and USA.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22089486-energy-distribution-measurement-narrow-band-ultrashort-ray-beams-via-edge-filters-subtraction','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22089486-energy-distribution-measurement-narrow-band-ultrashort-ray-beams-via-edge-filters-subtraction"><span><span class="hlt">Energy</span> distribution measurement of narrow-<span class="hlt">band</span> ultrashort x-ray beams via K-edge filters subtraction</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>Cardarelli, Paolo; Di Domenico, Giovanni; Marziani, Michele</p> <p>2012-10-01</p> <p>The characterization of novel x-ray sources includes the measurement of the photon flux and the <span class="hlt">energy</span> distribution of the produced beam. The aim of BEATS2 experiment at the SPARC-LAB facility of the INFN National Laboratories of Frascati (Rome, Italy) is to investigate possible medical applications of an x-ray source based on Thomson relativistic back-scattering. This source is expected to produce a pulsed quasi-monochromatic x-ray beam with an instantaneous flux of 10{sup 20} ph/s in pulses 10 ps long and with an average <span class="hlt">energy</span> of about 20 <span class="hlt">keV</span>. A direct measurement of <span class="hlt">energy</span> distribution of this beam is very difficult withmore » traditional detectors because of the extremely high photon flux. In this paper, we present a method for the evaluation of the <span class="hlt">energy</span> distribution of quasi-monochromatic x-ray beams based on beam filtration with K-edge absorbing foils in the <span class="hlt">energy</span> range of interest (16-22 <span class="hlt">keV</span>). The technique was tested measuring the <span class="hlt">energy</span> distribution of an x-ray beam having a spectrum similar to the expected one (SPARC-LAB Thomson source) by using a tungsten anode x-ray tube properly filtered and powered. The <span class="hlt">energy</span> distribution obtained has been compared with the one measured with a HPGe detector showing very good agreement.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19830032989&hterms=black+knight&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3Dblack%2Bknight','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19830032989&hterms=black+knight&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3Dblack%2Bknight"><span>Observations of 12-200 <span class="hlt">keV</span> X-rays from GX 339-4</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Nolan, P. L.; Gruber, D. E.; Knight, F. K.; Matteson, J. L.; Peterson, L. E.; Levine, A. M.; Lewin, W. H. G.; Primini, F. A.</p> <p>1982-01-01</p> <p>X-ray spectra of GX 339-4 measured on three occasions in 1977 and 1978 are presented. These are the first reported measurements above 10 <span class="hlt">keV</span>. The spectra can be described as the superposition of a soft component, which is dominant below about 20 <span class="hlt">keV</span>, and a hard component at higher <span class="hlt">energy</span>. Simultaneous measurements at lower <span class="hlt">energy</span> show that the soft component vanished during the observation in early 1978. The behavior of these two components is similar to that of the spectrum of Cygnus X-1; this reinforces the previously noted resemblance in rapid X-ray variability.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1347667-high-energy-gt-kev-ray-conversion-efficiency-measurement-arc-laser-national-ignition-facility','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1347667-high-energy-gt-kev-ray-conversion-efficiency-measurement-arc-laser-national-ignition-facility"><span>High-<span class="hlt">energy</span> (> 70 <span class="hlt">KeV</span>) x-ray conversion efficiency measurement on the ARC laser at the National Ignition Facility</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Chen, Hui; Hermann, M. R.; Kalantar, D. H.; ...</p> <p>2017-03-16</p> <p>Here, the Advanced Radiographic Capability (ARC) laser system at the National Ignition Facility (NIF) is designed to ultimately provide eight beamlets with a pulse duration adjustable from 1 to 30 ps, and <span class="hlt">energies</span> up to 1.5 kJ per beamlet. Currently, four beamlets have been commissioned. In the first set of 6 commissioning target experiments, the individual beamlets were fired onto gold foil targets with <span class="hlt">energy</span> up to 1 kJ per beamlet at 20–30 ps pulse length. The x-ray <span class="hlt">energy</span> distribution and pulse duration were measured, yielding <span class="hlt">energy</span> conversion efficiencies of 4–9 × 10 –4 for x-rays with <span class="hlt">energies</span> greater thanmore » 70 <span class="hlt">keV</span>. With greater than 3 J of such x-rays, ARC provides a high-precision x-ray backlighting capability for upcoming inertial confinement fusion and high-<span class="hlt">energy</span>-density physics experiments on NIF.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018PhRvC..97e4311C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PhRvC..97e4311C"><span>Rotational <span class="hlt">band</span> on a three-quasineutron isomer in 127Xe</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Chakraborty, S.; Sharma, H. P.; Tiwary, S. S.; Majumder, C.; Banerjee, P.; Ganguly, S.; Rai, S.; Pragati, Modi, Swati; Arumugam, P.; Mayank, Kumar, S.; Palit, R.; Kumar, A.; Bhattacharjee, S. S.; Singh, R. P.; Muralithar, S.</p> <p>2018-05-01</p> <p>Excited states in 127Xe were populated via 122Sn(9Be,4 n γ ) fusion-evaporation reaction at Ebeam=48 MeV. A rotational <span class="hlt">band</span> above an isomeric state at 2730.3 <span class="hlt">keV</span> has been established. Jπ=23 /2+ has been assigned unambiguously to the bandhead. Structural features, configuration, and reduced hindrance factor per degree of K -forbiddenness (fν) have been discussed in the context of neighboring nuclei. Half-lives of 7 /2+ (t1 /2=37 ±1 ns) and 23 /2+ (t1 /2=28 ±1 ns) isomeric states have also been estimated and found to be consistent with the earlier reported values. The experimentally observed <span class="hlt">energy</span> levels of the rotational <span class="hlt">band</span> have been reproduced successfully by theoretical calculations carried out using the modified particle rotor model (MPRM) with β2˜0.22 and 15∘≲γ ≲22∘ .</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19720013142','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19720013142"><span>Observation of Pulsed Gamma Rays with <span class="hlt">Energies</span> Greater than 250 <span class="hlt">keV</span> from the Crab Nebula Pulsar NP 0532. Ph.D. Thesis</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Orwig, L. E.</p> <p>1971-01-01</p> <p>Data from a balloon flight experiment using an Na I scintillation spectrometer were analyzed for gamma ray pulsation. The payload was carried to a nominal atmospheric depth of 3.5 g/cm2. A superposed epoch analysis was performed on a 12,000 second portion of the data spanning a total time interval of 16,700 seconds at float altitude. A positive pulsed contribution was observed at the expected apparent frequency of NP 0532 having the typical double pulse structure. The results indicate a time-averaged pulse photon flux of 0.00144 + or - 0.00057 photons/sq cm/sec in the <span class="hlt">energy</span> interval from 250 <span class="hlt">keV</span> to 2.3 MeV. This represents a time-averaged pulsed power of 0.000649 + or - 0.000257 keV/sq cm/sec/keV. The ratio of interpulse to main pulse was found to be 2.4 + or - 1.9. The analysis indicates a positive photon flux from 250 <span class="hlt">keV</span> to 725 <span class="hlt">keV</span> of 0.00120 + or - 0.00052 photons/sq cm/sec.</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 <span class="hlt">keV</span> 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('http://adsabs.harvard.edu/abs/2015AGUFMSM41F2559M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFMSM41F2559M"><span>Observations of Multi-<span class="hlt">band</span> Structures in Double Star TC-1 PEACE Electron and HIA Ion Data</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mohan Narasimhan, K.; Fazakerley, A. N.; Grimald, S.; Dandouras, I. S.; Mihaljcic, B.; Kistler, L. M.; Owen, C. J.</p> <p>2015-12-01</p> <p>Several authors have reported inner magnetosphere observations of proton distributions confined to narrow <span class="hlt">energy</span> <span class="hlt">bands</span> in the range 1 - 25 <span class="hlt">keV</span> (Smith and Hoffman (1974), etc). These structures have been described as "nose structures", with reference to their appearance in <span class="hlt">energy</span>-time spectrograms and are also known as "<span class="hlt">bands</span>" if they occur for extended periods of time. Multi-nose structures have been observed if 2 or more noses appear at the same time (Vallat et al., 2007). Gaps between "noses" (or "<span class="hlt">bands</span>") have been explained in terms of the competing corotation, convection and magnetic gradient drifts. Charge exchange losses in slow drift paths for steady state scenarios and the role of substorm injections have also been considered (Li et al., 2000; Ebihara et al., 2004). We analyse observations of electron and ion multi-<span class="hlt">band</span> structures frequently seen in Double-Star TC1 PEACE and HIA data. We present results from statistical surveys conducted using data from the duration of the mission. Furthermore, using a combination of both statistics and simulations, we test previous theories as to possible formation mechanisms and explore other possible explanations.</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/1988NIMPB..33..619J','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1988NIMPB..33..619J"><span>Surface analysis by means of high resolution <span class="hlt">energy</span> loss spectroscopy of 180° elastic scattered protons in the 100 <span class="hlt">keV</span> regime</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Jun-ichi, Kanasaki; Noriaki, Matsunami; Noriaki, Itoh; Tomoki, Oku; Kensin, Kitoh; Masahiko, Aoki; Koji, Matsuda</p> <p>1988-06-01</p> <p>The design and computer simulation of the performance of a new ion-beam surface analyzer has been presented. The analyzer has the capability of analyzing the <span class="hlt">energy</span> of ions incident at 100 <span class="hlt">keV</span> and scattered by 180° at surfaces with a resolution of 5 eV. The analyzer consists of an ion source, an accelerating-decelerating tube and a multichannel analyzer. Computer simulation of the <span class="hlt">energy</span> spectra of ions scattered from GaAs is reported.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012IJMPE..2150080H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012IJMPE..2150080H"><span>SEARCH FOR TWO-PHONON OCTUPOLE VIBRATIONAL <span class="hlt">BANDS</span> IN 88, 89, 92, 93, 94, 96Sr AND 95, 96, 97, 98Zr</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hwang, J. K.; Hamilton, J. H.; Ramayya, A. V.; Brewer, N. T.; Wang, E. H.; Luo, Y. X.; Zhu, S. J.</p> <p>2012-09-01</p> <p>Several new gamma transitions were identified in 94Sr, 93Sr, 92Sr, 96Zr and 97Zr from the spontaneous fission of 252Cf. Excited states in 88, 89, 92, 94, 96Sr and 95, 96, 97, 98Zr were reanalyzed and reorganized to propose the new two-phonon octupole vibrational states and <span class="hlt">bands</span>. The spin and parity of 6+ are assigned to a 4034.5 <span class="hlt">keV</span> state in 94Sr and 3576.4 <span class="hlt">keV</span> state in 98Zr. These states are proposed as the two-phonon octupole vibrational states along with the 6+ states at 3483.4 <span class="hlt">keV</span> in 96Zr, at 3786.0 <span class="hlt">keV</span> in 92Sr and 3604.2 <span class="hlt">keV</span> in 96Sr. The positive parity <span class="hlt">bands</span> in 88, 94, 96Sr and 96, 98Zr are the first two-phonon octupole vibrational <span class="hlt">bands</span> based on a 6+ state assigned in spherical nuclei. It is thought that in 94, 96Sr and 96, 98Zr a 3- octupole vibrational phonon is weakly coupled to an one-phonon octupole vibrational <span class="hlt">band</span> to make the two-phonon octupole vibrational <span class="hlt">band</span>. Also, the high spin states of odd-A95Zr and 97Zr are interpreted to be generated by the neutron 2d5/2 hole and neutron 1g7/2 particle, respectively, weakly coupled to one- and two-phonon octupole vibrational <span class="hlt">bands</span> of 96Zr. The high spin states of odd-A87Sr are interpreted to be caused by the neutron 1g9/2 hole weakly coupled to 3- and 5- states of 88Sr. New one- and two-POV <span class="hlt">bands</span> in 95, 97Zr and 87, 89Sr are proposed, for the first time, in the present work.</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('http://adsabs.harvard.edu/abs/2016PhRvA..93e2711W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016PhRvA..93e2711W"><span>Cross sections for ionization of tetrahydrofuran by protons at <span class="hlt">energies</span> between 300 and 3000 <span class="hlt">keV</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wang, Mingjie; Rudek, Benedikt; Bennett, Daniel; de Vera, Pablo; Bug, Marion; Buhr, Ticia; Baek, Woon Yong; Hilgers, Gerhard; Rabus, Hans</p> <p>2016-05-01</p> <p>Double-differential cross sections for ionization of tetrahydrofuran by protons with <span class="hlt">energies</span> from 300 to 3000 <span class="hlt">keV</span> were measured at the Physikalisch-Technische Bundesanstalt ion accelerator facility. The electrons emitted at angles between 15∘ and 150∘ relative to the ion-beam direction were detected with an electrostatic hemispherical electron spectrometer. Single-differential and total ionization cross sections have been derived by integration. The experimental results are compared to the semiempirical Hansen-Kocbach-Stolterfoht model as well as to the recently reported method based on the dielectric formalism. The comparison to the latter showed good agreement with experimental data in a broad range of emission angles and <span class="hlt">energies</span> of secondary electrons. The scaling property of ionization cross sections for tetrahydrofuran was also investigated. Compared to molecules of different size, the ionization cross sections of tetrahydrofuran were found to scale with the number of valence electrons at large impact parameters.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19790062013&hterms=palestine&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3Dpalestine','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19790062013&hterms=palestine&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3Dpalestine"><span>A possible line feature at 73 <span class="hlt">keV</span> from the Crab Nebula</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Ling, J. C.; Mahoney, W. A.; Willett, J. B.; Jacobson, A. S.</p> <p>1979-01-01</p> <p>Evidence is reported for a possible line feature at 73 <span class="hlt">keV</span> from the Crab Nebula. The experiment was conducted with a balloon-borne high-resolution gamma-ray spectrometer on June 10, 1974, over Palestine, Texas. The intensity and the width of the line derived from the fitting of these data are approximately 0.0038 photon per (sq cm-sec) and less than 4.9 <span class="hlt">keV</span> FWHM, respectively. The line is superposed on a power-law continuum of 11.2 E to the -2.16 photons per (sq cm-keV) in the <span class="hlt">energy</span> range from 53 to 300 <span class="hlt">keV</span>, which is consistent with other measurements of the Crab Nebula spectrum.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19980232021','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19980232021"><span>A Broad-<span class="hlt">band</span> Spectral and Timing Study of the X-Ray Binary System Centaurus X-3</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Audley, Michael Damian</p> <p>1998-01-01</p> <p>This dissertation describes a multi-mission investigation of the high mass X-ray binary pulsar Centaurus X-3. Cen X-3 was observed with the Broad <span class="hlt">Band</span> X-Ray Telescope (BBXRT) in December 1990. This was the first high-resolution solid state X-ray spectrometer to cover the iron K fluorescence region. The Fe K emission feature was resolved into two components for the first time. A broad 6.7 <span class="hlt">keV</span> feature was found to be a blend of lines from Fe XXI-Fe XXVI with <span class="hlt">energies</span> ranging from 6.6 to 6.9 <span class="hlt">keV</span> due to photoionization of the companion's stellar wind. A narrow line at 6.4 <span class="hlt">keV</span> due to fluorescence of iron in relatively low ionization states was also found. The quasi-periodic oscillations (QPO) at about 40 mHz were used to estimate the surface magnetic field of Cen X-3 as approx. 2.6 x 10(exp 12) G and to predict that there should be a cyclotron scattering resonance absorption feature (CSRF) near 30 <span class="hlt">keV</span>. In order to further resolve the iron line complex and to investigate the pulse-phase dependence of the iron line intensities, Cen X-3 was observed with the Advanced Satellite for Cosmology and Astrophysics (ASCA). Using ASCA's state-of-the-art non-dispersive X-ray spectrometers the 6.4 <span class="hlt">keV</span> fluorescent iron line was found to be pulsing while the intensities of the 6.7 and 6.9 <span class="hlt">keV</span> recombination lines do not vary with pulse phase. This confirms that the 6.4 <span class="hlt">keV</span> line is due to reflection by relatively neutral matter close to the neutron star while the recombination lines originate in the extended stellar wind. The continuum spectrum was found to be modified by reflection from matter close to the neutron star. Observations with the EXOSAT GSPC were used to search for a CSRF. The EXOSAT spectra were consistent with the presence of a CSRF but an unambiguous detection was not possible because of a lack of sensitivity at <span class="hlt">energies</span> higher than the cyclotron <span class="hlt">energy</span>. Cen X-3 was then observed with the Rossi X-Ray Timing Explorer (RXTE) and evidence for a CSRF at 25.1 +/- 0.3 <span class="hlt">keV</span> was</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22591198-comparison-martian-meteorites-earth-composition-study-effective-atomic-numbers-energy-range-kev-gev','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22591198-comparison-martian-meteorites-earth-composition-study-effective-atomic-numbers-energy-range-kev-gev"><span>Comparison of Martian meteorites with earth composition: Study of effective atomic numbers in the <span class="hlt">energy</span> range 1 <span class="hlt">keV</span>-100 GeV</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>Ün, Adem, E-mail: ademun25@yahoo.com; Han, İbrahim, E-mail: ibrahimhan25@hotmail.com; Ün, Mümine, E-mail: mun@agri.edu.tr</p> <p>2016-04-18</p> <p>Effective atomic (Z{sub eff}) and electron numbers (N{sub eff}) for 24 Martian meteorites have been determined in the <span class="hlt">energy</span> range from 1 <span class="hlt">keV</span> to 100 GeV and also for sixteen significant <span class="hlt">energies</span> of commonly used radioactive sources. The values of Z{sub eff} and N{sub eff} for all sample were obtained from the DirectZeff program. The obtained results for Martian meteorites have been compared with the results for Earth composition and similarities or differences also evaluated.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018IJT....39...74O','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018IJT....39...74O"><span>Photocarrier Radiometry for Non-contact Evaluation of Monocrystalline Silicon Solar Cell Under Low-<span class="hlt">Energy</span> (< 200 <span class="hlt">keV</span>) Proton Irradiation</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Oliullah, Md.; Liu, J. Y.; Song, P.; Wang, Y.</p> <p>2018-06-01</p> <p>A three-layer theoretical model is developed for the characterization of the electronic transport properties (lifetime τ, diffusion coefficient D, and surface recombination velocity s) with energetic particle irradiation on solar cells using non-contact photocarrier radiometry. Monte Carlo (MC) simulation is carried out to obtain the depth profiles of the proton irradiation layer at different low <span class="hlt">energies</span> (< 200 <span class="hlt">keV</span>). The monocrystalline silicon (c-Si) solar cells are investigated under different low-<span class="hlt">energy</span> proton irradiation, and the carrier transport parameters of the three layers are obtained by best-fitting of the experimental results. The results show that the low-<span class="hlt">energy</span> protons have little influence on the transport parameters of the non-irradiated layer, but high influences on both of the p and n-region irradiation layers which are consisted of MC simulation.</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 additional terms in the k .p expansion of the Hamiltonian.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19820049534&hterms=598&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D80%26Ntt%3D598','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19820049534&hterms=598&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D80%26Ntt%3D598"><span>Venus nightside ionosphere - A model with <span class="hlt">KeV</span> electron impact ionization</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Kumar, S.</p> <p>1982-01-01</p> <p>The impact of <span class="hlt">keV</span> electrons is proposed as the strongest source of ionization in a full-up Venus nightside ionosphere model for the equatorial midnight region. The electron impacts lead to a peak ion density of 100,000/cu cm, which was observed by the PV-OIMS experiment on several occasions. In addition, the observed altitude profiles of CO2(+), O(+), O2(+), H(+), and H2(+) can be reproduced by the model on condition that the available <span class="hlt">keV</span> electron flux is approximated by a reasonable extrapolation from fluxes observed at lower <span class="hlt">energies</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20040055313&hterms=atom&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3Datom','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20040055313&hterms=atom&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3Datom"><span>Remote Sensing of Icy Galilean Moon Surface and Atmospheric Composition Using Low <span class="hlt">Energy</span> (1 eV-4 <span class="hlt">keV</span>) Neutral Atom Imaging</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Collier, M. R.; Sittler, E.; Chornay, D.; Cooper, J. F.; Coplan, M.; Johnson, R. E.</p> <p>2004-01-01</p> <p>We describe a low <span class="hlt">energy</span> neutral atom imager suitable for composition measurements Europa and other icy Galilean moons in the Jovian magnetosphere. This instrument employs conversion surface technology and is sensitive to either neutrals converted to negative ions, neutrals converted to positive ions and the positive ions themselves depending on the power supply. On a mission such as the Jupiter Icy Moons Orbiter (JIMO), two back-to-back sensors would be flown with separate power supplies fitted to the neutral atom and iodneutral atom sides. This will allow both remote imaging of 1 eV < E < 4 <span class="hlt">keV</span> neutrals from icy moon surfaces and atmospheres, and in situ measurements of ions at similar <span class="hlt">energies</span> in the moon ionospheres and Jovian magnetospheric plasma. The instrument provides composition measurements of the neutrals and ions that enter the spectrometer with a mass resolution dependent on the time-of-flight subsystem and capable of resolving molecules. The lower <span class="hlt">energy</span> neutrals, up to tens of eV, arise from atoms and molecules sputtered off the moon surfaces and out of the moon atmospheres by impacts of more energetic (<span class="hlt">keV</span> to MeV) ions from the magnetosphere. Direct Simulation Monte Carlo (DSMC) models are used to convert measured neutral abundances to compositional distributions of primary and trace species in the sputtered surfaces and atmospheres. The escaping neutrals can also be detected as ions after photo- or plasma-ionization and pickup. Higher <span class="hlt">energy</span>, <span class="hlt">keV</span> neutrals come from charge exchange of magnetospheric ions in the moon atmospheres and provide information on atmospheric structure. At the jovicentric orbits of the icy moons the presence of toroidal gas clouds, as detected at Europa's orbit, provide M e r opportunities to analyze both the composition of neutrals and ions originating from the moon surfaces, and the characteristics of magnetospheric ions interacting with neutral cloud material. Charge exchange of low <span class="hlt">energy</span> ions near the moons, and</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/2016SPIE.9968E..03V','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016SPIE.9968E..03V"><span>High <span class="hlt">energy</span> resolution plastic scintillator</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>van Loef, Edgar V.; Feng, Patrick; Markosyan, Gary; Shirwadkar, Urmila; Doty, Patrick; Shah, Kanai S.</p> <p>2016-09-01</p> <p>In this paper we present results on a novel tin-loaded plastic scintillator. We will show that this particular plastic scintillator has a light output similar to that of BGO, a fast scintillation decay (< 10 ns), exhibits good neutron/gamma PSD with a Figure-of-Merit of 1.3 at 2.5 MeVee cut-off <span class="hlt">energy</span>, and excellent <span class="hlt">energy</span> resolution of about 12% (FWHM) at 662 <span class="hlt">keV</span>. Under X-ray excitation, the radioluminescence spectrum exhibits a broad <span class="hlt">band</span> between 350 and 500 nm peaking at 420 nm which is well-matched to bialkali photomultiplier tubes and UV-enhanced photodiodes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25085141','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25085141"><span>Moiré deflectometry using the Talbot-Lau interferometer as refraction diagnostic for high <span class="hlt">energy</span> density plasmas at <span class="hlt">energies</span> below 10 <span class="hlt">keV</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Valdivia, M P; Stutman, D; Finkenthal, M</p> <p>2014-07-01</p> <p>The highly localized density gradients expected in High <span class="hlt">Energy</span> Density (HED) plasma experiments can be characterized by x-ray phase-contrast imaging in addition to conventional attenuation radiography. Moiré deflectometry using the Talbot-Lau grating interferometer setup is an attractive HED diagnostic due to its high sensitivity to refraction induced phase shifts. We report on the adaptation of such a system for operation in the sub-10 <span class="hlt">keV</span> range by using a combination of free standing and ultrathin Talbot gratings. This new x-ray <span class="hlt">energy</span> explored matches well the current x-ray backlighters used for HED experiments, while also enhancing phase effects at lower electron densities. We studied the performance of the high magnification, low <span class="hlt">energy</span> Talbot-Lau interferometer, for single image phase retrieval using Moiré fringe deflectometry. Our laboratory and simulation studies indicate that such a device is able to retrieve object electron densities from phase shift measurements. Using laboratory x-ray sources from 7 to 15 μm size we obtained accurate simultaneous measurements of refraction and attenuation for both sharp and mild electron density gradients.</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://hdl.handle.net/2060/20070031884','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20070031884"><span>The Morphology of the X-ray Emission above 2 <span class="hlt">keV</span> from Jupiter's Aurorae</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Elsner, R.; Branduardi-Raymont, G.; Galand, M.; Grodent, D.; Waite, J. H.; Cravens, T.; Ford, P.</p> <p>2007-01-01</p> <p>The discovery in XMM-Newton X-ray data of X-ray emission above 2 <span class="hlt">keV</span> from Jupiter's aurorae has led us to reexamine the Chandra ACIS-S observations taken in Feb 2003. Chandra's superior spatial resolution has revealed that the auroral X-rays with E > 2 <span class="hlt">keV</span> are emitted from the periphery of the region emitting those with E < 1 <span class="hlt">keV</span>. We are presently exploring the relationship of this morphology to that of the FUV emission from the main auroral oval and the polar cap. The low <span class="hlt">energy</span> emission has previously been established as due to charge exchange between energetic precipitating ions of oxygen and either sulfur or carbon. It seems likely to us that the higher <span class="hlt">energy</span> emission is due to precipitation of energetic electrons, possibly the same population of electrons responsible for the FUV emission. We discuss our analysis and interpretation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19720048772&hterms=causes+absorption+window&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3DWhat%2Bcauses%2Babsorption%2Bwindow','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19720048772&hterms=causes+absorption+window&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3DWhat%2Bcauses%2Babsorption%2Bwindow"><span>A sealed titanium window proportional counter for the detection of .5-<span class="hlt">keV</span> X rays.</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Mcclintock, J. E.; Levine, A.; Rappaport, S.</p> <p>1972-01-01</p> <p>A sealed Ti window proportional counter sensitive to X radiation in the <span class="hlt">energy</span> range 0.35-0.45 <span class="hlt">keV</span> and above 1.5 <span class="hlt">keV</span> is described. Measurements of the Ti mass absorption coefficients and a graphical summary of the literature values are presented. For a proportional counter with a 930 microgram/sq cm (2.1-micron) Ti window, the peak efficiency at 0.45 <span class="hlt">keV</span> is found to lie between 4.6% and 7.1%. An application in X-ray astronomy involving a rocket observation of Sco X-1 is discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20150008976&hterms=spectroscopy&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dspectroscopy','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20150008976&hterms=spectroscopy&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dspectroscopy"><span>Broad-<span class="hlt">Band</span> Spectroscopy of Hercules X-1 with Suzaku</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Asami, Fumi; Enoto, Teruaki; Iwakiri, Wataru; Yamada, Shin'ya; Tamagawa, Toru; Mihara, Tatehiro; Nagase, Fumiaki</p> <p>2014-01-01</p> <p>Hercules X-1 was observed with Suzaku in the main-on state from 2005 to 2010. The 0.4- 100 <span class="hlt">keV</span> wide-<span class="hlt">band</span> spectra obtained in four observations showed a broad hump around 4-9 <span class="hlt">keV</span> in addition to narrow Fe lines at 6.4 and 6.7 <span class="hlt">keV</span>. The hump was seen in all the four observations regardless of the selection of the continuum models. Thus it is considered a stable and intrinsic spectral feature in Her X-1. The broad hump lacked a sharp structure like an absorption edge. Thus it was represented by two different spectral models: an ionized partial covering or an additional broad line at 6.5 <span class="hlt">keV</span>. The former required a persistently existing ionized absorber, whose origin was unclear. In the latter case, the Gaussian fitting of the 6.5-<span class="hlt">keV</span> line needs a large width of sigma = 1.0-1.5 <span class="hlt">keV</span> and a large equivalent width of 400-900 eV. If the broad line originates from Fe fluorescence of accreting matter, its large width may be explained by the Doppler broadening in the accretion flow. However, the large equivalent width may be inconsistent with a simple accretion geometry.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5524379','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5524379"><span>Calculations of Electron Inelastic Mean Free Paths. XI. Data for Liquid Water for <span class="hlt">Energies</span> from 50 eV to 30 <span class="hlt">keV</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>Shinotsuka, H.; Da, B.; Tanuma, S.; Yoshikawa, H.; Powell, C. J.; Penn, D. R.</p> <p>2017-01-01</p> <p>We calculated electron inelastic mean free paths (IMFPs) for liquid water from its optical <span class="hlt">energy</span>-loss function (ELF) for electron <span class="hlt">energies</span> from 50 eV to 30 <span class="hlt">keV</span>. These calculations were made with the relativistic full Penn algorithm (FPA) that has been used for previous IMFP and electron stopping-power calculations for many elemental solids. We also calculated IMFPs of water with three additional algorithms: the relativistic single-pole approximation (SPA), the relativistic simplified SPA, and the relativistic extended Mermin method. These calculations were made using the same optical ELF in order to assess any differences of the IMFPs arising from choice of the algorithm. We found good agreement among the IMFPs from the four algorithms for <span class="hlt">energies</span> over 300 eV. For <span class="hlt">energies</span> less than 100 eV, however, large differences became apparent. IMFPs from the relativistic TPP-2M equation for predicting IMFPs were in good agreement with IMFPs from the four algorithms for <span class="hlt">energies</span> between 300 eV and 30 <span class="hlt">keV</span> but there was poorer agreement for lower <span class="hlt">energies</span>. We calculated values of the static structure factor as a function of momentum transfer from the FPA. The resulting values were in good agreement with results from first-principles calculations and with inelastic X-ray scattering spectroscopy experiments. We made comparisons of our IMFPs with earlier calculations from authors who had used different algorithms and different ELF data sets. IMFP differences could then be analyzed in terms of the algorithms and the data sets. Finally, we compared our IMFPs with measurements of IMFPs and of a related quantity, the effective attenuation length (EAL). There were large variations in the measured IMFPs and EALs (as well as their dependence on electron <span class="hlt">energy</span>). Further measurements are therefore required to establish consistent data sets and for more detailed comparisons with calculated IMFPs. PMID:28751796</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28751796','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28751796"><span>Calculations of Electron Inelastic Mean Free Paths. XI. Data for Liquid Water for <span class="hlt">Energies</span> from 50 eV to 30 <span class="hlt">keV</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Shinotsuka, H; Da, B; Tanuma, S; Yoshikawa, H; Powell, C J; Penn, D R</p> <p>2017-04-01</p> <p>We calculated electron inelastic mean free paths (IMFPs) for liquid water from its optical <span class="hlt">energy</span>-loss function (ELF) for electron <span class="hlt">energies</span> from 50 eV to 30 <span class="hlt">keV</span>. These calculations were made with the relativistic full Penn algorithm (FPA) that has been used for previous IMFP and electron stopping-power calculations for many elemental solids. We also calculated IMFPs of water with three additional algorithms: the relativistic single-pole approximation (SPA), the relativistic simplified SPA, and the relativistic extended Mermin method. These calculations were made using the same optical ELF in order to assess any differences of the IMFPs arising from choice of the algorithm. We found good agreement among the IMFPs from the four algorithms for <span class="hlt">energies</span> over 300 eV. For <span class="hlt">energies</span> less than 100 eV, however, large differences became apparent. IMFPs from the relativistic TPP-2M equation for predicting IMFPs were in good agreement with IMFPs from the four algorithms for <span class="hlt">energies</span> between 300 eV and 30 <span class="hlt">keV</span> but there was poorer agreement for lower <span class="hlt">energies</span>. We calculated values of the static structure factor as a function of momentum transfer from the FPA. The resulting values were in good agreement with results from first-principles calculations and with inelastic X-ray scattering spectroscopy experiments. We made comparisons of our IMFPs with earlier calculations from authors who had used different algorithms and different ELF data sets. IMFP differences could then be analyzed in terms of the algorithms and the data sets. Finally, we compared our IMFPs with measurements of IMFPs and of a related quantity, the effective attenuation length (EAL). There were large variations in the measured IMFPs and EALs (as well as their dependence on electron <span class="hlt">energy</span>). Further measurements are therefore required to establish consistent data sets and for more detailed comparisons with calculated IMFPs.</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/1993SPIE.1736...36B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1993SPIE.1736...36B"><span>Absolute detection efficiency of a microchannel plate detector to X rays in the 1-100 <span class="hlt">KeV</span> <span class="hlt">energy</span> range</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Burginyon, Gary A.; Jacoby, Barry A.; Wobser, James K.; Ernst, Richard; Ancheta, Dione S.; Tirsell, Kenneth G.</p> <p>1993-02-01</p> <p>There is little information in the literature on the performance of working micro-channel plate (MCP) detectors at high x-ray <span class="hlt">energies</span>. We have measured the absolute efficiency of a microchannel-plate-intensified, subnanosecond, one dimensional imaging x-ray detector developed at LLNL in the 1 to 100 <span class="hlt">keV</span> range and at 1.25 MeV. The detector consists of a gold photocathode deposited on the front surface of the MCP (optimized for Ni K(subscript (alpha) ) x rays) to convert x rays to electrons, an MCP to amplify the electrons, and a fast In:CdS phosphor that converts the electron's kinetic <span class="hlt">energy</span> to light. The phosphor is coated on a fiber-optic faceplate to transmit the light out of the vacuum system. Electrostatic focusing electrodes compress the electron current out of the MCP in one dimension while preserving spatial resolution in the other. The calibration geometry, dictated by a recent experiment, required grazing incidence x rays (15.6 degree(s)) onto the MCP detector in order to maximize deliverable current. The experiment also used a second detector made up of 0.071 in. thick BC422 plastic scintillator material from the Bicron Corporation. We compare the absolute efficiencies of these two detectors in units of optical W/cm(superscript 2) into 4 (pi) per x ray W/cm(superscript 2) incident. At 7.47 <span class="hlt">keV</span> and 900 volts MCP bias, the MCP detector delivers approximately 1400 times more light than the scintillator detector.</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> </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('http://adsabs.harvard.edu/abs/1986SPIE..689..155G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1986SPIE..689..155G"><span>Facilities and Techniques for X-Ray Diagnostic Calibration in the 100-eV to 100-<span class="hlt">keV</span> <span class="hlt">Energy</span> Range</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gaines, J. L.; Wittmayer, F. J.</p> <p>1986-08-01</p> <p>The Lawrence Livermore National Laboratory (LLNL) has been a pioneer in the field of x-ray diagnostic calibration for more than 20 years. We have built steady state x-ray sources capable of supplying fluorescent lines of high spectral purity in the 100-eV to 100-<span class="hlt">keV</span> <span class="hlt">energy</span> range, and these sources have been used in the calibration of x-ray detectors, mirrors, crystals, filters, and film. This paper discusses our calibration philosophy and techniques, and describes some of our x-ray sources. Examples of actual calibration data are presented as well.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22590920-solar-wind-kev-electrons-quiet-times','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22590920-solar-wind-kev-electrons-quiet-times"><span>Solar wind ∼0.1-1.5 <span class="hlt">keV</span> electrons at quiet times</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>Tao, Jiawei; Wang, Linghua, E-mail: wanglhwang@gmail.com; Zong, Qiugang</p> <p>2016-03-25</p> <p>We present a statistical survey of the <span class="hlt">energy</span> spectrum of solar wind suprathermal (∼0.1-1.5 <span class="hlt">keV</span>) electrons measured by the WIND 3-D Plasma & Energetic Particle (3DP) instrument at 1 AU during quiet times at the minimum and maximum of solar cycles 23 and 24. Firstly, we separate strahl (beaming) electrons and halo (isotropic) electrons based on their features in pitch angle distributions. Secondly, we fit the observed <span class="hlt">energy</span> spectrum of both the strahl and halo electrons at ∼0.1-1.5 <span class="hlt">keV</span> to a Kappa distribution function with an index κ, effective temperature T{sub eff} and density n{sub 0}. We also integrate themore » the measurements over ∼0.1-1.5 <span class="hlt">keV</span> to obtain the average electron <span class="hlt">energy</span> E{sub avg} of the strahl and halo. We find a strong positive correlation between κ and T{sub eff} for both the strahl and halo, possibly reflecting the nature of the generation of these suprathermal electrons. Among the 245 selected samples, ∼68% have the halo κ smaller than the strahl κ, while ∼50% have the halo E{sub h} larger than the strahl E{sub s}.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2008EPJD...48..383M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2008EPJD...48..383M"><span>Transmission of ˜ 10 <span class="hlt">keV</span> electron beams through thin ceramic foils: Measurements and Monte Carlo simulations of electron <span class="hlt">energy</span> distribution functions</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Morozov, A.; Heindl, T.; Skrobol, C.; Wieser, J.; Krücken, R.; Ulrich, A.</p> <p>2008-07-01</p> <p>Electron beams with particle <span class="hlt">energy</span> of ~10 <span class="hlt">keV</span> were sent through 300 nm thick ceramic (Si3N4 + SiO2) foils and the resulting electron <span class="hlt">energy</span> distribution functions were recorded using a retarding grid technique. The results are compared with Monte Carlo simulations performed with two publicly available packages, Geant4 and Casino v2.42. It is demonstrated that Geant4, unlike Casino, provides electron <span class="hlt">energy</span> distribution functions very similar to the experimental distributions. Both simulation packages provide a quite precise average <span class="hlt">energy</span> of transmitted electrons: we demonstrate that the maximum uncertainty of the calculated values of the average <span class="hlt">energy</span> is 6% for Geant4 and 8% for Casino, taking into account all systematic uncertainties and the discrepancies in the experimental and simulated data.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25430208','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25430208"><span>Absolute calibration of the Agfa Structurix series films at <span class="hlt">energies</span> between 2.7 and 6.2 <span class="hlt">keV</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Lanier, N E; Cowan, J S</p> <p>2014-11-01</p> <p>Although photo-emulsion technology is many decades old, x-ray film still remains a key asset for diagnosing hydrodynamic features in High-<span class="hlt">Energy</span> Density (HED) experiments. For decades, the preferred option had been Kodak's direct exposure film. After its discontinuance in 2004, the push to find alternatives began. In many situations, the Agfa Structurix series offers the most favorable substitute, but being new to the HED community, its characterization was lacking. To remedy this, recent experiments, conducted at Brookhaven's National Synchrotron Light Source, provide absolute, monochromatic calibration data for the Agfa Structurix series films at K-shell backlighter <span class="hlt">energies</span> between 2.7 and 6.2 <span class="hlt">keV</span>. Absolute response curves are presented for Agfa D8, D7, D4, D4sc, D3, and D2. Moreover, the transmission of each film type is also measured.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22308827-moire-deflectometry-using-talbot-lau-interferometer-refraction-diagnostic-high-energy-density-plasmas-energies-below-kev','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22308827-moire-deflectometry-using-talbot-lau-interferometer-refraction-diagnostic-high-energy-density-plasmas-energies-below-kev"><span>Moiré deflectometry using the Talbot-Lau interferometer as refraction diagnostic for High <span class="hlt">Energy</span> Density plasmas at <span class="hlt">energies</span> below 10 <span class="hlt">keV</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>Valdivia, M. P.; Stutman, D.; Finkenthal, M.</p> <p>2014-07-15</p> <p>The highly localized density gradients expected in High <span class="hlt">Energy</span> Density (HED) plasma experiments can be characterized by x-ray phase-contrast imaging in addition to conventional attenuation radiography. Moiré deflectometry using the Talbot-Lau grating interferometer setup is an attractive HED diagnostic due to its high sensitivity to refraction induced phase shifts. We report on the adaptation of such a system for operation in the sub-10 <span class="hlt">keV</span> range by using a combination of free standing and ultrathin Talbot gratings. This new x-ray <span class="hlt">energy</span> explored matches well the current x-ray backlighters used for HED experiments, while also enhancing phase effects at lower electron densities.more » We studied the performance of the high magnification, low <span class="hlt">energy</span> Talbot-Lau interferometer, for single image phase retrieval using Moiré fringe deflectometry. Our laboratory and simulation studies indicate that such a device is able to retrieve object electron densities from phase shift measurements. Using laboratory x-ray sources from 7 to 15 μm size we obtained accurate simultaneous measurements of refraction and attenuation for both sharp and mild electron density gradients.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016IAUS..320..112A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016IAUS..320..112A"><span>Thermal characteristics of a B8.3 flare observed on July 04, 2009</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Awasthi, Arun Kumar; Sylwester, Barbara; Sylwester, Janusz; Jain, Rajmal</p> <p></p> <p>We explore the temporal evolution of flare plasma parameters including temperature (T) - differential emission measure (DEM) relationship by analyzing high spectral and temporal cadence of X-ray emission in 1.6-8.0 <span class="hlt">keV</span> <span class="hlt">energy</span> <span class="hlt">band</span>, recorded by SphinX (Polish) and Solar X-ray Spectrometer (SOXS; Indian) instruments, during a B8.3 flare which occurred on July 04, 2009. SphinX records X-ray emission in 1.2-15.0 <span class="hlt">keV</span> <span class="hlt">energy</span> <span class="hlt">band</span> with the temporal and spectral cadence as good as 6 μs and 0.4 <span class="hlt">keV</span>, respectively. On the other hand, SOXS provides X-ray observations in 4-25 <span class="hlt">keV</span> <span class="hlt">energy</span> <span class="hlt">band</span> with the temporal and spectral resolution of 3 s and 0.7 <span class="hlt">keV</span>, respectively. We derive the thermal plasma parameters during impulsive phase of the flare employing well-established Withbroe-Sylwester DEM inversion algorithm.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011MNRAS.418.2367B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011MNRAS.418.2367B"><span>The high-<span class="hlt">energy</span> view of the broad-line radio galaxy 3C 111</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ballo, L.; Braito, V.; Reeves, J. N.; Sambruna, R. M.; Tombesi, F.</p> <p>2011-12-01</p> <p>We present the analysis of Suzaku and XMM-Newton observations of the broad-line radio galaxy (BLRG) 3C 111. Its high-<span class="hlt">energy</span> emission shows variability, a harder continuum with respect to the radio-quiet active galactic nucleus population, and weak reflection features. Suzaku found the source in a minimum flux level; a comparison with the XMM-Newton data implies an increase of a factor of 2.5 in the 0.5-10 <span class="hlt">keV</span> flux, in the 6 months separating the two observations. The iron K complex is detected in both data sets, with rather low equivalent width(s). The intensity of the iron K complex does not respond to the change in continuum flux. An ultrafast, high-ionization outflowing gas is clearly detected in the Suzaku/X-ray Imaging Spectrometer data; the absorber is most likely unstable. Indeed, during the XMM-Newton observation, which was 6 months after, the absorber was not detected. No clear rollover in the hard X-ray emission is detected, probably due to the emergence of the jet as a dominant component in the hard X-ray <span class="hlt">band</span>, as suggested by the detection above ˜100 <span class="hlt">keV</span> with the GSO onboard Suzaku, although the present data do not allow us to firmly constrain the relative contribution of the different components. The fluxes observed by the γ-ray satellites CGRO and Fermi would be compatible with the putative jet component if peaking at <span class="hlt">energies</span> E˜ 100 MeV. In the X-ray <span class="hlt">band</span>, the jet contribution to the continuum starts to be significant only above 10 <span class="hlt">keV</span>. If the detection of the jet component in 3C 111 is confirmed, then its relative importance in the X-ray <span class="hlt">energy</span> <span class="hlt">band</span> could explain the different observed properties in the high-<span class="hlt">energy</span> emission of BLRGs, which are otherwise similar in their other multiwavelength properties. Comparison between X-ray and γ-ray data taken at different epochs suggests that the strong variability observed for 3C 111 is probably driven by a change in the primary continuum.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFMSH51A2467Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFMSH51A2467Y"><span>The Strongest 40 <span class="hlt">keV</span> Electron Acceleration By ICME-driven Shocks At 1 AU</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yang, L.; Wang, L.; Li, G.; Wimmer-Schweingruber, R. F.; He, J.; Tu, C. Y.; Bale, S. D.</p> <p>2017-12-01</p> <p>Here we present a comprehensive case study of the in situ electron acceleration at the two ICME-driven shocks observed by WIND/3DP on February 11, 2000 and July 22, 2004. For the 11 February 2000 shock (the 22 July 2004 shock), the shocked electrons in the downstream show significant flux enhancements over the ambient solar wind electrons at <span class="hlt">energies</span> up to 40 <span class="hlt">keV</span> (66 <span class="hlt">keV</span>) with a 6.0 times (1.9 times) ehancment at 40 <span class="hlt">keV</span>, the strongest among all the quasi-perpendicular (quasi-parallel) ICME-driven shocks observed by the WIND spacecraft at 1 AU from 1995 through 2014. We find that in both shocks, the shocked electron fluxes at 0.5-40 <span class="hlt">keV</span> fit well to a double power-law spectrum, J ˜ E-β, bending up at ˜2 <span class="hlt">keV</span>. In the downstream, these shocked electrons show stronger fluxes in the anti-sunward direction, but their enhancement over the ambient fluxes peaks near 90° pitch angle (PA). For the 11 February 2000 shock, the electron spectral index, β, appears to not vary with the electron PA, while for the 22 July 2004 shock, β roughly decreases from the anti-sunward PA direction to the sunward PA direction. All of these spectral indexes are strongly larger than the theoretical prediction of diffusive shock acceleration. At <span class="hlt">energies</span> above (below) 2 <span class="hlt">keV</span>, however, the shocked electron β is similar to the solar wind superhalo (halo) electrons observed at quiet times. These results suggest that the electron acceleration at the ICME-driven shocks at 1 AU may favor the shock drift acceleration, and the superthermal electrons accelerated by the interplanetary shocks may contribute to the formation of the halo and superhalo electron populations in the solar wind.</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/2017RaPC..130..343A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017RaPC..130..343A"><span>Photon interaction study of organic nonlinear optical materials in the <span class="hlt">energy</span> range 122-1330 <span class="hlt">keV</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Awasarmol, Vishal V.; Gaikwad, Dhammajyot K.; Raut, Siddheshwar D.; Pawar, Pravina P.</p> <p>2017-01-01</p> <p>In the present study, the mass attenuation coefficient (μm) of six organic nonlinear optical materials has been calculated in the <span class="hlt">energy</span> range 122-1330 <span class="hlt">keV</span> and compared with the obtained values from the WinXCOM program. It is found that there is a good agreement between theoretical and experimental values (<3%). The linear attenuation coefficients (μ) total atomic cross section (σt, a), and total electronic cross section (σt, el) have also been calculated from the obtained μm values and their variations with photon <span class="hlt">energy</span> have been plotted. From the present work, it is observed that the variation of obtained values of μm, μ, σt, a, and σt, el strongly depends on the photon <span class="hlt">energy</span> and decreases or increases due to chemical composition and density of the sample. All the samples have been studied extensively using transmission method with a view to utilize the material for radiation dosimetry. Investigated samples are good material for radiation dosimetry due their low effective atomic number. The mass attenuation coefficient (μm), linear attenuation coefficients (μ), total atomic cross section (σt, a), total electronic cross section (σt, el), effective atomic numbers (Zeff), molar extinction coefficient (ε), mass <span class="hlt">energy</span> absorption coefficient (μen/ρ) and effective atomic <span class="hlt">energy</span> absorption cross section (σa, en) of all sample materials have been carried out and transmission curves have been plotted. The transmission curve shows that the variation of all sample materials decreases with increasing photon <span class="hlt">energy</span>.</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('https://ntrs.nasa.gov/search.jsp?R=20100031229&hterms=ev&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dev','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20100031229&hterms=ev&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dev"><span>The Swift/Fermi GRB 080928 from 1 eV to 150 <span class="hlt">keV</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Sonbas, Eda; Rossi, A.; Schulze, S.; Klose, S.; Kann, D. A.; Ferrero, P.; NicuesaGuelbenzu, A.; Rau, A.; Kruehler, T.; Greiner, J.; <a style="text-decoration: none; " href="javascript:void(0); " onClick="displayelement('author_20100031229'); toggleEditAbsImage('author_20100031229_show'); toggleEditAbsImage('author_20100031229_hide'); "> <img style="display:inline; width:12px; height:12px; " src="images/arrow-up.gif" width="12" height="12" border="0" alt="hide" id="author_20100031229_show"> <img style="width:12px; height:12px; display:none; " src="images/arrow-down.gif" width="12" height="12" border="0" alt="hide" id="author_20100031229_hide"></p> <p>2010-01-01</p> <p>We present the results of a comprehensive study of the Gamma-Ray Burst 080928 and of its afterglow. GRB 08092 was a long burst detected by Swift/BAT and Fermi/GBM, It is one of the exceptional cases where optical emission was already detected when the GRB itself was still radiating in the gamma-ray <span class="hlt">band</span>. for nearly 100 seconds simultaneous optical X-ray and gamma-ray data provide a coverage of the spectral <span class="hlt">energy</span> distribution of the transient source from about 1 eV to 150 <span class="hlt">keV</span>. Here we analyze the prompt emission, constrain its spectral propertIes. and set lower limits on the initial Lorentz factor of the relativistic outflow, In particular. we show that the SED during the main prompt emission phase is in agreement with synchrotron radiation. We construct the optical/near-infrared light curve and the spectral <span class="hlt">energy</span> distribution based on Swift/UVOT. ROTSE-Illa (Australia) and GROND (La Silla) data and compare it to the X-ray light curve retrieved from the Swift/XRT repository. We show that its bumpy shape can be modeled by multiple <span class="hlt">energy</span> injections into the forward shock. Furthermore, we provide evidence that the temporal and spectral evolution of the first strong flare seen in the early X-ray light curve can be explained by large-angle emission. Finally, we report on the results of our search for the GRB host galaxy, for which only a deep upper limit can be provided.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1994NIMPB..93..505K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1994NIMPB..93..505K"><span>Low noise scintillation detectors with a P-47 thin layer screen for electrons of several <span class="hlt">keV</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kajcsos, Zs.; Meisel, W.; Griesbach, P.; Gütlich, P.; Sauer, Ch.; Kurz, R.; Hildebrand, K.; Albrecht, R.; Ligtenberg, M. A. C.</p> <p>1994-09-01</p> <p>The applicability of a low-noise scintillation detector (ScD) for the registration of electrons of several <span class="hlt">keV</span> <span class="hlt">energy</span> has been studied employing photomultipliers (PM) of different types and sizes. With the application of a sedimented P-47 scintillation screen, the values of the low-<span class="hlt">energy</span> sensitivity limit and those of the light conversion coefficient were determined as about 2.7-4.7 <span class="hlt">keV</span> and 2.8-6.6 photoelectrons/keV, respectively, for the set of PM's (Philips-Valvo XP 2020, Philips-Valvo XP 2052, Philips-Valvo XP 2972, EMI 9124a) studied. It is concluded that such scintillation detectors might be used advantageously as electron counters in the range of E > 5 <span class="hlt">keV</span>. Applications below this kinetic <span class="hlt">energy</span> value are also feasible when applying a floating acceleration of several kV to the ScD — a voltage much lower than the values required for Everhart-Thornley detectors.</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> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AIPC.1833b0083A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AIPC.1833b0083A"><span>Calculation of absorption parameters for selected narcotic drugs in the <span class="hlt">energy</span> range from 1 <span class="hlt">keV</span> to 100 GeV</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Akman, Ferdi; Kaçal, Mustafa Recep; Akdemir, Fatma; Araz, Aslı; Turhan, Mehmet Fatih; Durak, Rıdvan</p> <p>2017-04-01</p> <p>The total mass attenuation coefficients (μ/ρ), total molecular (σt,m), atomic (σt,a) and electronic (σt,e) cross sections, effective atomic numbers (Zeff) and electron density (NE) were computed in the wide <span class="hlt">energy</span> region from 1 <span class="hlt">keV</span> to 100 GeV for the selected narcotic drugs such as morphine, heroin, cocaine, ecstasy and cannabis. The changes of μ/ρ, σt,m, σt,a, σt,e, Zeff and NE with photon <span class="hlt">energy</span> for total photon interaction shows the dominance of different interaction process in different <span class="hlt">energy</span> regions. The variations of μ/ρ, σt,m, σt,a, σt,e, Zeff and NE depend on the atom number, photon <span class="hlt">energy</span> and chemical composition of narcotic drugs. Also, these parameters change with number of elements, the range of atomic numbers in narcotic drugs and total molecular weight. These data can be useful in the field of forensic sciences and medical diagnostic.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1998SPIE.3301...27V','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1998SPIE.3301...27V"><span>6 x 6-cm fully depleted pn-junction CCD for high-resolution spectroscopy in the 0.1- to 15-<span class="hlt">keV</span> photon <span class="hlt">energy</span> range</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>von Zanthier, Christoph; Holl, Peter; Kemmer, Josef; Lechner, Peter; Maier, B.; Soltau, Heike; Stoetter, R.; Braeuninger, Heinrich W.; Dennerl, Konrad; Haberl, Frank; Hartmann, R.; Hartner, Gisela D.; Hippmann, H.; Kastelic, E.; Kink, W.; Krause, N.; Meidinger, Norbert; Metzner, G.; Pfeffermann, Elmar; Popp, M.; Reppin, Claus; Stoetter, Diana; Strueder, Lothar; Truemper, Joachim; Weber, U.; Carathanassis, D.; Engelhard, S.; Gebhart, Th.; Hauff, D.; Lutz, G.; Richter, R. H.; Seitz, H.; Solc, P.; Bihler, Edgar; Boettcher, H.; Kendziorra, Eckhard; Kraemer, J.; Pflueger, Bernhard; Staubert, Ruediger</p> <p>1998-04-01</p> <p>The concept and performance of the fully depleted pn- junction CCD system, developed for the European XMM- and the German ABRIXAS-satellite missions for soft x-ray imaging and spectroscopy in the 0.1 <span class="hlt">keV</span> to 15 <span class="hlt">keV</span> photon range, is presented. The 58 mm X 60 mm large pn-CCD array uses pn- junctions for registers and for the backside instead of MOS registers. This concept naturally allows to fully deplete the detector volume to make it an efficient detector to photons with <span class="hlt">energies</span> up to 15 <span class="hlt">keV</span>. For high detection efficiency in the soft x-ray region down to 100 eV, an ultrathin pn-CCD backside deadlayer has been realized. Each pn-CCD-channel is equipped with an on-chip JFET amplifier which, in combination with the CAMEX-amplifier and multiplexing chip, facilitates parallel readout with a pixel read rate of 3 MHz and an electronic noise floor of ENC < e-. With the complete parallel readout, very fast pn-CCD readout modi can be implemented in the system which allow for high resolution photon spectroscopy of even the brightest x-ray sources in the sky.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016APS..DPPNO8008S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016APS..DPPNO8008S"><span>A 7.2 <span class="hlt">keV</span> spherical crystal backlighter system for Sandia's Z Pulsed Power Facility</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Schollmeier, M.; Knapp, P. F.; Ampleford, D. J.; Loisel, G. P.; Robertson, G.; Shores, J. E.; Smith, I. C.; Speas, C. S.; Porter, J. L.; McBride, R. D.</p> <p>2016-10-01</p> <p>Many experiments on Sandia's Z facility, a 30 MA, 100 ns rise-time, pulsed-power driver, use a monochromatic Quartz crystal imaging backlighter system at 1.865 <span class="hlt">keV</span> (Si Heα) or 6.151 <span class="hlt">keV</span> (Mn Heα) x-ray <span class="hlt">energy</span> to radiograph an imploding liner (cylindrical tube) or wire array. The x-ray source is generated by the Z-Beamlet Laser (ZBL), which provides up to 4.5 kJ at 527 nm during a 6 ns window. Radiographs of an imploding thick-walled Beryllium liner at a convergence ratio of about 20 [CR =Rin . (0) /Rin . (t) ] were too opaque to identify the inner surface of the liner with high confidence, demonstrating the need for a higher-<span class="hlt">energy</span> x-ray backlighter between 6 and 10 <span class="hlt">keV</span>. We present the design, test and first application of a Ge (335) spherical crystal x-ray backlighter system using the 7.242 <span class="hlt">keV</span> Co Heα resonance line. The system operates at an almost identical Bragg angle as the existing 1.865 and 6.151 <span class="hlt">keV</span> backlighters, enhancing our capabilities such as two-color, two-frame radiography, without changing detector shielding hardware. SAND No: SAND2016-6724 A. Sandia National Laboratories is a multi-program laboratory managed and operated by Sandia Corp., a wholly owned subsidiary of Lockheed Martin Corp., for the U.S. DoE NNSA under contract DE-AC04-94AL85000.</p> </li> <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('https://www.ncbi.nlm.nih.gov/pubmed/28798216','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28798216"><span>Dual-<span class="hlt">Energy</span> CT in Enhancing Subdural Effusions that Masquerade as Subdural Hematomas: Diagnosis with Virtual High-Monochromatic (190-<span class="hlt">keV</span>) Images.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Bodanapally, U K; Dreizin, D; Issa, G; Archer-Arroyo, K L; Sudini, K; Fleiter, T R</p> <p>2017-10-01</p> <p>Extravasation of iodinated contrast into subdural space following contrast-enhanced radiographic studies results in hyperdense subdural effusions, which can be mistaken as acute subdural hematomas on follow-up noncontrast head CTs. Our aim was to identify the factors associated with contrast-enhancing subdural effusion, characterize diffusion and washout kinetics of iodine in enhancing subdural effusion, and assess the utility of dual-<span class="hlt">energy</span> CT in differentiating enhancing subdural effusion from subdural hematoma. We retrospectively analyzed follow-up head dual-<span class="hlt">energy</span> CT studies in 423 patients with polytrauma who had undergone contrast-enhanced whole-body CT. Twenty-four patients with enhancing subdural effusion composed the study group, and 24 randomly selected patients with subdural hematoma were enrolled in the comparison group. Postprocessing with syngo.via was performed to determine the diffusion and washout kinetics of iodine. The sensitivity and specificity of dual-<span class="hlt">energy</span> CT for the diagnosis of enhancing subdural effusion were determined with 120-kV, virtual monochromatic <span class="hlt">energy</span> (190-<span class="hlt">keV</span>) and virtual noncontrast images. Patients with enhancing subdural effusion were significantly older (mean, 69 years; 95% CI, 60-78 years; P < .001) and had a higher incidence of intracranial hemorrhage ( P = .001). Peak iodine concentration in enhancing subdural effusions was reached within the first 8 hours of contrast administration with a mean of 0.98 mg/mL (95% CI, 0.81-1.13 mg/mL), and complete washout was achieved at 38 hours. For the presence of a hyperdense subdural collection on 120-kV images with a loss of hyperattenuation on 190-<span class="hlt">keV</span> and virtual noncontrast images, when considered as a true-positive for enhancing subdural effusion, the sensitivity was 100% (95% CI, 85.75%-100%) and the specificity was 91.67% (95% CI, 73%-99%). Dual-<span class="hlt">energy</span> CT has a high sensitivity and specificity in differentiating enhancing subdural effusion from subdural hematoma. Hence, dual-<span class="hlt">energy</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_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('https://www.osti.gov/biblio/22494523-study-parameters-scanning-system-kev-electron-accelerator','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22494523-study-parameters-scanning-system-kev-electron-accelerator"><span>Study on the parameters of the scanning system for the 300 <span class="hlt">keV</span> electron accelerator</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>Leo, K. W.; Chulan, R. M., E-mail: leo@nm.gov.my; Hashim, S. A.</p> <p>2016-01-22</p> <p>This paper describes the method to identify the magnetic coil parameters of the scanning system. This locally designed low <span class="hlt">energy</span> electron accelerator with the present <span class="hlt">energy</span> of 140 <span class="hlt">keV</span> will be upgraded to 300 <span class="hlt">keV</span>. In this accelerator, scanning system is required to deflect the energetic electron beam across a titanium foil in vertical and horizontal direction. The excitation current of the magnetic coil is determined by the <span class="hlt">energy</span> of the electron beam. Therefore, the magnetic coil parameters must be identified to ensure the matching of the beam <span class="hlt">energy</span> and excitation coil current. As the result, the essential parameters ofmore » the effective lengths for X-axis and Y-axis have been found as 0.1198 m and 0.1134 m and the required excitation coil currents which is dependenton the electron beam <span class="hlt">energies</span> have be identified.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/21239846','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21239846"><span>Comparison of GATE/GEANT4 with EGSnrc and MCNP for electron dose calculations at <span class="hlt">energies</span> between 15 <span class="hlt">keV</span> and 20 MeV.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Maigne, L; Perrot, Y; Schaart, D R; Donnarieix, D; Breton, V</p> <p>2011-02-07</p> <p>The GATE Monte Carlo simulation platform based on the GEANT4 toolkit has come into widespread use for simulating positron emission tomography (PET) and single photon emission computed tomography (SPECT) imaging devices. Here, we explore its use for calculating electron dose distributions in water. Mono-energetic electron dose point kernels and pencil beam kernels in water are calculated for different <span class="hlt">energies</span> between 15 <span class="hlt">keV</span> and 20 MeV by means of GATE 6.0, which makes use of the GEANT4 version 9.2 Standard Electromagnetic Physics Package. The results are compared to the well-validated codes EGSnrc and MCNP4C. It is shown that recent improvements made to the GEANT4/GATE software result in significantly better agreement with the other codes. We furthermore illustrate several issues of general interest to GATE and GEANT4 users who wish to perform accurate simulations involving electrons. Provided that the electron step size is sufficiently restricted, GATE 6.0 and EGSnrc dose point kernels are shown to agree to within less than 3% of the maximum dose between 50 <span class="hlt">keV</span> and 4 MeV, while pencil beam kernels are found to agree to within less than 4% of the maximum dose between 15 <span class="hlt">keV</span> and 20 MeV.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22531481-determination-neutron-spectra-within-energy-kev-mev-means-reactor-dosimetry','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22531481-determination-neutron-spectra-within-energy-kev-mev-means-reactor-dosimetry"><span>Determination of neutron spectra within the <span class="hlt">energy</span> of 1 <span class="hlt">keV</span> to 1 MeV by means of reactor dosimetry</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>Sergeyeva, Victoria; Destouches, Christophe; Lyoussi, Abdallah</p> <p>2015-07-01</p> <p>The standard procedure for neutron reactor dosimetry is based on neutron irradiation of a target and its post-irradiation analysis by Gamma and/or X-ray spectrometry. Nowadays, the neutron spectra can be easily characterized for thermal and fast <span class="hlt">energies</span> (respectively 0.025 eV and >1 MeV). In this work we propose a new target and an innovating post-irradiation technique of analysis in order to detect the neutron spectra within the <span class="hlt">energy</span> of 1 <span class="hlt">keV</span> to 1 MeV. This article will present the calculations performed for the selection of a suitable nuclear reaction and isotope, the results predicted by simulations, the irradiation campaign thatmore » is proposed and the post-irradiation technique of analysis. (authors)« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19830056692&hterms=Cross+fit&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3DCross%2Bfit','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19830056692&hterms=Cross+fit&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3DCross%2Bfit"><span>Interstellar photoelectric absorption cross sections, 0.03-10 <span class="hlt">keV</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Morrison, R.; Mccammon, D.</p> <p>1983-01-01</p> <p>An effective absorption cross section per hydrogen atom has been calculated as a function of <span class="hlt">energy</span> in the 0.03-10 <span class="hlt">keV</span> range using the most recent atomic cross section and cosmic abundance data. Coefficients of a piecewise polynomial fit to the numerical results are given to allow convenient application in automated calculations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015PhPl...22d3114H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015PhPl...22d3114H"><span>Developing a bright 17 <span class="hlt">keV</span> x-ray source for probing high-<span class="hlt">energy</span>-density states of matter at high spatial resolution</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Huntington, C. M.; Park, H.-S.; Maddox, B. R.; Barrios, M. A.; Benedetti, R.; Braun, D. G.; Hohenberger, M.; Landen, O. L.; Regan, S. P.; Wehrenberg, C. E.; Remington, B. A.</p> <p>2015-04-01</p> <p>A set of experiments were performed on the National Ignition Facility (NIF) to develop and optimize a bright, 17 <span class="hlt">keV</span> x-ray backlighter probe using laser-irradiated Nb foils. High-resolution one-dimensional imaging was achieved using a 15 μm wide slit in a Ta substrate to aperture the Nb Heα x-rays onto an open-aperture, time integrated camera. To optimize the x-ray source for imaging applications, the effect of laser pulse shape and spatial profile on the target was investigated. Two laser pulse shapes were used—a "prepulse" shape that included a 3 ns, low-intensity laser foot preceding the high-<span class="hlt">energy</span> 2 ns square main laser drive, and a pulse without the laser foot. The laser spatial profile was varied by the use of continuous phase plates (CPPs) on a pair of shots compared to beams at best focus, without CPPs. A comprehensive set of common diagnostics allowed for a direct comparison of imaging resolution, total x-ray conversion efficiency, and x-ray spectrum between shots. The use of CPPs was seen to reduce the high-<span class="hlt">energy</span> tail of the x-ray spectrum, whereas the laser pulse shape had little effect on the high-<span class="hlt">energy</span> tail. The measured imaging resolution was comparably high for all combinations of laser parameters, but a higher x-ray flux was achieved without phase plates. This increased flux was the result of smaller laser spot sizes, which allowed us to arrange the laser focal spots from multiple beams and produce an x-ray source which was more localized behind the slit aperture. Our experiments are a first demonstration of point-projection geometry imaging at NIF at the <span class="hlt">energies</span> (>10 <span class="hlt">keV</span>) necessary for imaging denser, higher-Z targets than have previously been investigated.</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/2004NIMPB.225..198S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2004NIMPB.225..198S"><span>Measurements of Rayleigh, Compton and resonant Raman scattering cross-sections for 59.536 <span class="hlt">keV</span> γ-rays</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Singh, Prem; Mehta, D.; Singh, N.; Puri, S.; Shahi, J. S.</p> <p>2004-09-01</p> <p>The K-L and K-M resonant Raman scattering (RRS) cross-sections have been measured for the first time at the 59.536 <span class="hlt">keV</span> photon <span class="hlt">energy</span> in the 70Yb ( BK=61.332 <span class="hlt">keV</span>), 71Lu ( BK=63.316 <span class="hlt">keV</span>) and 72Hf ( BK=65.345 <span class="hlt">keV</span>) elements; BK being the K-shell binding <span class="hlt">energy</span>. The K-L and K-M RRS measurements have been performed at the 59° and 133° angles, respectively, to avoid interference of the Compton-scatter peak. The Rayleigh and Compton scattering cross-sections for the 59.536 <span class="hlt">keV</span> γ-rays have also been measured at both the angles in the atomic region 1⩽ Z⩽92. Measurements were performed using the reflection-mode geometrical arrangements involving the 241Am radioisotope as photon source and planar Si(Li) and HPGe detectors. Ratios of the K-M and K-L RRS cross-sections in Yb, Lu and Hf are in general lower than that of the fluorescent Kβ 1,3,5 (K-M) and Kα (K-L) X-ray transition probabilities. Theoretical Rayleigh scattering cross-sections based on the modified form-factors (MFs) corrected for the anomalous scattering factors (ASFs) and the S-matrix calculations are on an average ˜15% and ˜6% higher, respectively, at the 133° angle and exhibit good agreement with the measured data at the 59° angle. Larger deviations ˜30% and ˜20%, respectively, are observed at the 133° angle for the 64Gd, 66Dy, 67Ho and 70Yb elements having the K-shell binding <span class="hlt">energy</span> in vicinity of the incident photon <span class="hlt">energy</span>. The measured Compton scattering cross-sections are in general agreement with those calculated using the Klein-Nishina cross-sections and the incoherent scattering function.</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('http://adsabs.harvard.edu/abs/2016AcASn..57..139W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AcASn..57..139W"><span>Periodicity Analysis of X-ray Light Curves of SS 433</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wang, J. Y.; Lu, X. L.; Zhao, Q. W.; Dong, D. Q.; Lao, B. Q.; Lu, Y.; Wei, Y. H.; Wu, X. C.; An, T.</p> <p>2016-03-01</p> <p>SS 433 is the only X-ray binary to date that was detected to have a pair of well-collimated jets, and its orbital period, super orbital period, and nutation period were all detected at the same time. The study on the periodic X-ray variabilities is helpful for understanding its dynamic process of the central engine and the correlation with other <span class="hlt">bands</span>. In the present paper, two time series analysis techniques, Lomb-Scargle periodogram and weighted wavelet Z-transform, are employed to search for the periodicities from the Swift/BAT (Burst Alert Telescope)(15--50 <span class="hlt">keV</span>) and RXTE/ASM (Rossi X-Ray Timing Explorer/All-Sky Monitor)(1.5--3, 3--5 and 5--12 <span class="hlt">keV</span>) light curves of SS 433, and the Monte Carlo simulation is performed. For the 15--50 <span class="hlt">keV</span> <span class="hlt">energy</span> <span class="hlt">band</span>, five significant periodic signals are detected, which are P_1(˜6.29 d), P_2 (˜6.54 d), P_3 (˜13.08 d), P_4 (˜81.50 d), and P_5 (˜162.30 d). For the 3--5 and 5--12 <span class="hlt">keV</span> <span class="hlt">energy</span> <span class="hlt">bands</span>, periodic signals P_3 (˜13 d) and P_5 (˜162 d) are detected in both <span class="hlt">energy</span> <span class="hlt">bands</span>. However, for the 1.5--3 <span class="hlt">keV</span> <span class="hlt">energy</span> <span class="hlt">band</span>, no significant periodic signal is detected. P_5 has the strongest periodic signal in the power spectrum for all the <span class="hlt">energy</span> <span class="hlt">bands</span> of 3--5, 5--12, and 15--50 <span class="hlt">keV</span>, and it is consistent with that obtained by previous study in optical <span class="hlt">band</span>. Further, due to the existence of relativistic radio jets, the X-ray and optical <span class="hlt">band</span> variability of P_5 (˜162 d) is probably related to the precession of the relativistic jets. High coherence between X-ray and optical light curves may also imply that the X-ray and optical emissions are of the same physical origin. P_3 shows a good agreement with the orbital period (˜13.07 d) first obtained by previous study, and P_2 and P_4 are the high frequency harmonic components of P_3 and P_5, respectively. P_1 is detected from the power spectrum of 15--50 <span class="hlt">keV</span> <span class="hlt">energy</span> <span class="hlt">band</span> only, and it is consistent with the systematic nutation period. As the power of <span class="hlt">energy</span> <span class="hlt">band</span> decreases (from hard X-ray to</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/6921407-imaging-observations-sn1987a-gamma-ray-energies','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/6921407-imaging-observations-sn1987a-gamma-ray-energies"><span>Imaging observations of SN1987A at gamma-ray <span class="hlt">energies</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>Cook, W.R.; Palmer, D.M.; Prince, T.A.</p> <p>1988-09-25</p> <p>The Caltech imaging ..gamma..-ray telescope was launched by balloon from Alice Springs, NT, Australia for observations of SN1987A during the period 18.60--18.87 November 1987 UT. The preliminary results presented here are derived from 8200 seconds of instrument livetime on the supernova and 2500 seconds on the Crab Nebula and pulsar at a float altitude of 37 km. We have obtained the first images of the SN1987A region at ..gamma..-ray <span class="hlt">energies</span> confirming that the bulk of the ..gamma..-ray emission comes from the supernova and not from LMC X-1. A count excess is detected between 300 and 1300 <span class="hlt">keV</span> from the directionmore » of the supernova, one third of which comes from <span class="hlt">energy</span> <span class="hlt">bands</span> of width 80 and 92 <span class="hlt">keV</span> centered on 847 and 1238 <span class="hlt">keV</span>, respectively. The excess can be interpreted as a line photon flux plus scattered photon continuum from the radioactive decay of /sup 56/Co synthesized in the supernova explosion. We compare our data to recent predictions and find it to be consistent with models invoking moderate mixing of core material into the envelope.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1988AIPC..170...60C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1988AIPC..170...60C"><span>Imaging observations of SN1987A at gamma-ray <span class="hlt">energies</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Cook, W. R.; Palmer, D. M.; Prince, T. A.; Schindler, S. M.; Starr, C. H.; Stone, E. C.</p> <p>1988-09-01</p> <p>The Caltech imaging γ-ray telescope was launched by balloon from Alice Springs, NT, Australia for observations of SN1987A during the period 18.60-18.87 November 1987 UT. The preliminary results presented here are derived from 8200 seconds of instrument livetime on the supernova and 2500 seconds on the Crab Nebula and pulsar at a float altitude of 37 km. We have obtained the first images of the SN1987A region at γ-ray <span class="hlt">energies</span> confirming that the bulk of the γ-ray emission comes from the supernova and not from LMC X-1. A count excess is detected between 300 and 1300 <span class="hlt">keV</span> from the direction of the supernova, one third of which comes from <span class="hlt">energy</span> <span class="hlt">bands</span> of width 80 and 92 <span class="hlt">keV</span> centered on 847 and 1238 <span class="hlt">keV</span>, respectively. The excess can be interpreted as a line photon flux plus scattered photon continuum from the radioactive decay of 56Co synthesized in the supernova explosion. We compare our data to recent predictions and find it to be consistent with models invoking moderate mixing of core material into the envelope.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018PhRvC..97d1601M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PhRvC..97d1601M"><span>Unifying measurement of 239Pu(n ,γ ) in the <span class="hlt">keV</span> to MeV <span class="hlt">energy</span> regime</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mosby, S.; Bredeweg, T. A.; Couture, A.; Jandel, M.; Kawano, T.; Ullmann, J. L.; Henderson, R. A.; Wu, C. Y.</p> <p>2018-04-01</p> <p>A single, unifying measurement of the 239Pu capture cross section from 1 <span class="hlt">keV</span> to 1.3 MeV has been performed for the first time using the Detector for Advanced Neutron Capture Experiments (DANCE) at the Los Alamos Neutron Science Center (LANSCE). The experimental method combines a prior experiment's characterization of prompt fission γ rays in conjunction with a fission tagging detector with a separate experiment using a thick 239Pu sample to extract the neutron capture cross section in ratio to 239Pu(n ,f ) . We have made new predictions of the capture cross section taking into account recent results for the M 1 scissors mode present in other actinides. The results show deviations from current evaluations which are 30% higher at the highest <span class="hlt">energies</span>, and will be used to improve calculations relevant for several applications.</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('http://adsabs.harvard.edu/abs/2015RaPC..107..199S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015RaPC..107..199S"><span>Electron density of Rhizophora spp. wood using Compton scattering technique at 15.77, 17.48 and 22.16 <span class="hlt">keV</span> XRF <span class="hlt">energies</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Shakhreet, B. Z.; Bauk, S.; Shukri, A.</p> <p>2015-02-01</p> <p>Compton (incoherently) scattered photons which are directly proportional to the electron density of the scatterer, have been employed in characterizing Rhizophora spp. as breast tissue equivalent. X-ray fluorescent scattered incoherently from Rhizophora spp. sample was measured using Si-PIN detector and three XRF <span class="hlt">energy</span> values 15.77, 17.48 and 22.16 <span class="hlt">keV</span>. This study is aimed at providing electron density information in support of the introduction of new tissue substitute materials for mammography phantoms.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19790049763&hterms=oso&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Doso','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19790049763&hterms=oso&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Doso"><span>The MIT/OSO 7 catalog of X-ray sources - Intensities, spectra, and long-term variability</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Markert, T. H.; Laird, F. N.; Clark, G. W.; Hearn, D. R.; Sprott, G. F.; Li, F. K.; Bradt, H. V.; Lewin, W. H. G.; Schnopper, H. W.; Winkler, P. F.</p> <p>1979-01-01</p> <p>This paper is a summary of the observations of the cosmic X-ray sky performed by the MIT 1-40-<span class="hlt">keV</span> X-ray detectors on OSO 7 between October 1971 and May 1973. Specifically, mean intensities or upper limits of all third Uhuru or OSO 7 cataloged sources (185 sources) in the 3-10-<span class="hlt">keV</span> range are computed. For those sources for which a statistically significant (greater than 20) intensity was found in the 3-10-<span class="hlt">keV</span> <span class="hlt">band</span> (138 sources), further intensity determinations were made in the 1-15-<span class="hlt">keV</span>, 1-6-<span class="hlt">keV</span>, and 15-40-<span class="hlt">keV</span> <span class="hlt">energy</span> <span class="hlt">bands</span>. Graphs and other simple techniques are provided to aid the user in converting the observed counting rates to convenient units and in determining spectral parameters. Long-term light curves (counting rates in one or more <span class="hlt">energy</span> <span class="hlt">bands</span> as a function of time) are plotted for 86 of the brighter sources.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018PhRvC..97f4603P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PhRvC..97f4603P"><span>Measurement and resonance analysis of the 33S(n ,α )30Si cross section at the CERN n_TOF facility in the <span class="hlt">energy</span> region from 10 to 300 <span class="hlt">keV</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Praena, J.; Sabaté-Gilarte, M.; Porras, I.; Quesada, J. M.; Altstadt, S.; Andrzejewski, J.; Audouin, L.; Bécares, V.; Barbagallo, M.; Bečvář, F.; Belloni, F.; Berthoumieux, E.; Billowes, J.; Boccone, V.; Bosnar, D.; Brugger, M.; Calviño, F.; Calviani, M.; Cano-Ott, D.; Carrapiço, C.; Cerutti, F.; Chiaveri, E.; Chin, M.; Colonna, N.; Cortés, G.; Cortés-Giraldo, M. A.; Diakaki, M.; Dietz, M.; Domingo-Pardo, C.; Dressler, R.; Durán, I.; Eleftheriadis, C.; Ferrari, A.; Fraval, K.; Furman, V.; Göbel, K.; Gómez-Hornillos, M. B.; Ganesan, S.; García, A. R.; Giubrone, G.; Gonçalves, I. F.; González-Romero, E.; Goverdovski, A.; Griesmayer, E.; Guerrero, C.; Gunsing, F.; Heftrich, T.; Hernández-Prieto, A.; Heyse, J.; Jenkins, D. G.; Jericha, E.; Käppeler, F.; Kadi, Y.; Karadimos, D.; Katabuchi, T.; Ketlerov, V.; Khryachkov, V.; Kivel, N.; Koehler, P.; Kokkoris, M.; Kroll, J.; Krtička, M.; Lampoudis, C.; Langer, C.; Leal-Cidoncha, E.; Lederer-Woods, C.; Leeb, H.; Leong, L. S.; Lerendegui-Marco, J.; Losito, R.; Mallick, A.; Manousos, A.; Marganiec, J.; Martínez, T.; Massimi, C.; Mastinu, P.; Mastromarco, M.; Mendoza, E.; Mengoni, A.; Milazzo, P. M.; Mingrone, F.; Mirea, M.; Mondelaers, W.; Paradela, C.; Pavlik, A.; Perkowski, J.; Plompen, A. J. M.; Rauscher, T.; Reifarth, R.; Riego-Perez, A.; Robles, M.; Rubbia, C.; Ryan, J. A.; Sarmento, R.; Saxena, A.; Schillebeeckx, P.; Schmidt, S.; Schumann, D.; Sedyshev, P.; Tagliente, G.; Tain, J. L.; Tarifeño-Saldivia, A.; Tarrío, D.; Tassan-Got, L.; Tsinganis, A.; Valenta, S.; Vannini, G.; Variale, V.; Vaz, P.; Ventura, A.; Vermeulen, M. J.; Vlachoudis, V.; Vlastou, R.; Wallner, A.; Ware, T.; Weigand, M.; Weiss, C.; Wright, T.; Žugec, P.; n TOF Collaboration</p> <p>2018-06-01</p> <p>The 33S(n ,α )30Si cross section has been measured at the neutron time-of-flight (n _TOF ) facility at CERN in the neutron <span class="hlt">energy</span> range from 10 to 300 <span class="hlt">keV</span> relative to the 10B(n ,α )7Li cross-section standard. Both reactions were measured simultaneously with a set of micromegas detectors. The flight path of 185 m has allowed us to obtain the cross section with high-<span class="hlt">energy</span> resolution. An accurate description of the resonances has been performed by means of the multilevel multichannel R -matrix code sammy. The results show a significantly higher area of the biggest resonance (13.45 <span class="hlt">keV</span>) than the unique high-resolution (n ,α ) measurement. The new parametrization of the 13.45-<span class="hlt">keV</span> resonance is similar to that of the unique transmission measurement. This resonance is a matter of research in neutron-capture therapy. The 33S(n ,α )30Si cross section has been studied in previous works because of its role in the production of 36S in stars, which is currently overproduced in stellar models compared to observations.</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.osti.gov/pages/biblio/1255852-energy-dependent-dynamics-kev-mev-electrons-inner-zone-outer-zone-slot-regions','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1255852-energy-dependent-dynamics-kev-mev-electrons-inner-zone-outer-zone-slot-regions"><span><span class="hlt">Energy</span>-dependent dynamics of <span class="hlt">keV</span> to MeV electrons in the inner zone, outer zone, and slot regions</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Reeves, Geoffrey D.; Friedel, Reiner H. W.; Larsen, Brian A.; ...</p> <p>2016-01-28</p> <p>Here, we present observations of the radiation belts from the Helium Oxygen Proton Electron and Magnetic Electron Ion Spectrometer particle detectors on the Van Allen Probes satellites that illustrate the <span class="hlt">energy</span> dependence and L shell dependence of radiation belt enhancements and decays. We survey events in 2013 and analyze an event on 1 March in more detail. The observations show the following: (a) at all L shells, lower <span class="hlt">energy</span> electrons are enhanced more often than higher <span class="hlt">energies</span>; (b) events that fill the slot region are more common at lower <span class="hlt">energies</span>; (c) enhancements of electrons in the inner zone are moremore » common at lower <span class="hlt">energies</span>; and (d) even when events do not fully fill the slot region, enhancements at lower <span class="hlt">energies</span> tend to extend to lower L shells than higher <span class="hlt">energies</span>. During enhancement events the outer zone extends to lower L shells at lower <span class="hlt">energies</span> while being confined to higher L shells at higher <span class="hlt">energies</span>. The inner zone shows the opposite with an outer boundary at higher L shells for lower <span class="hlt">energies</span>. Both boundaries are nearly straight in log(<span class="hlt">energy</span>) versus L shell space. At <span class="hlt">energies</span> below a few 100 <span class="hlt">keV</span>, radiation belt electron penetration through the slot region into the inner zone is commonplace, but the number and frequency of “slot filling” events decreases with increasing <span class="hlt">energy</span>. The inner zone is enhanced only at <span class="hlt">energies</span> that penetrate through the slot. <span class="hlt">Energy</span>- and L shell-dependent losses (that are consistent with whistler hiss interactions) return the belts to more quiescent conditions.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20140017786','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20140017786"><span>The NuSTAR Extragalactic Survey: A First Sensitive Look at the High-<span class="hlt">Energy</span> Cosmic X-Ray Background Population</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Alexander, D. M.; Stern, D.; DelMoro, A.; Lansbury, G. B.; Assef, R. J.; Aird, J.; Ajello, M.; Ballantyne, D. R.; Bauer, F. E.; Boggs, S. E.; <a style="text-decoration: none; " href="javascript:void(0); " onClick="displayelement('author_20140017786'); toggleEditAbsImage('author_20140017786_show'); toggleEditAbsImage('author_20140017786_hide'); "> <img style="display:inline; width:12px; height:12px; " src="images/arrow-up.gif" width="12" height="12" border="0" alt="hide" id="author_20140017786_show"> <img style="width:12px; height:12px; display:none; " src="images/arrow-down.gif" width="12" height="12" border="0" alt="hide" id="author_20140017786_hide"></p> <p>2013-01-01</p> <p>We report on the first 10 identifications of sources serendipitously detected by the Nuclear Spectroscopic Telescope Array (NuSTAR) to provide the first sensitive census of the cosmic X-ray background source population at approximately greater than 10 <span class="hlt">keV</span>. We find that these NuSTAR-detected sources are approximately 100 times fainter than those previously detected at approximately greater than 10 <span class="hlt">keV</span> and have a broad range in redshift and luminosity (z = 0.020-2.923 and L(sub 10-40 <span class="hlt">keV</span>) approximately equals 4 × 10(exp 41) - 5 × 10(exp 45) erg per second; the median redshift and luminosity are z approximately equal to 0.7 and L(sub 10-40 <span class="hlt">keV</span>) approximately equal to 3 × 10(exp 44) erg per second, respectively. We characterize these sources on the basis of broad-<span class="hlt">band</span> approximately equal to 0.5 - 32 <span class="hlt">keV</span> spectroscopy, optical spectroscopy, and broad-<span class="hlt">band</span> ultraviolet-to-mid-infrared spectral <span class="hlt">energy</span> distribution analyses. We find that the dominant source population is quasars with L(sub 10-40 <span class="hlt">keV</span>) greater than 10(exp 44) erg per second, of which approximately 50% are obscured with N(sub H) approximately greater than 10(exp 22) per square centimeters. However, none of the 10 NuSTAR sources are Compton thick (N(sub H) approximately greater than 10(exp 24) per square centimeters) and we place a 90% confidence upper limit on the fraction of Compton-thick quasars (L(sub 10-40 <span class="hlt">keV</span>) greater than 10(exp 44) erg per second) selected at approximately greater than 10 <span class="hlt">keV</span> of approximately less than 33% over the redshift range z = 0.5 - 1.1. We jointly fitted the rest-frame approximately equal to 10-40 <span class="hlt">keV</span> data for all of the non-beamed sources with L(sub 10-40 <span class="hlt">keV</span>) greater than 10(exp 43) erg per second to constrain the average strength of reflection; we find R less than 1.4 for gamma = 1.8, broadly consistent with that found for local active galactic nuclei (AGNs) observed at approximately greater than 10 <span class="hlt">keV</span>. We also constrain the host-galaxy masses and find a median stellar</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('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.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.ncbi.nlm.nih.gov/pubmed/25464176','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25464176"><span>Measurement of X-ray mass attenuation coefficients in biological and geological samples in the <span class="hlt">energy</span> range of 7-12<span class="hlt">keV</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Trunova, Valentina; Sidorina, Anna; Kriventsov, Vladimir</p> <p>2015-01-01</p> <p>Information about X-ray mass attenuation coefficients in different materials is necessary for accurate X-ray fluorescent analysis. The X-ray mass attenuation coefficients for <span class="hlt">energy</span> of 7-12<span class="hlt">keV</span> were measured in biological (Mussel and Oyster tissues, blood, hair, liver, and Cabbage leaves) and geological (Baikal sludge, soil, and Alaskite granite) samples. The measurements were carried out at the EXAFS Station of Siberian Synchrotron Radiation Center (VEPP-3). Obtained experimental mass attenuation coefficients were compared with theoretical values calculated for some samples. Copyright © 2014 Elsevier Ltd. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1455403-unifying-measurement-pu239-kev-mev-energy-regime','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1455403-unifying-measurement-pu239-kev-mev-energy-regime"><span>Unifying measurement of Pu 239 ( n , γ ) in the <span class="hlt">keV</span> to MeV <span class="hlt">energy</span> regime</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Mosby, S.; Bredeweg, T. A.; Couture, A.; ...</p> <p>2018-04-23</p> <p>A single, unifying measurement of the 239Pu capture cross section from 1 <span class="hlt">keV</span> to 1.3 MeV has been performed for the first time using the Detector for Advanced Neutron Capture Experiments (DANCE) at the Los Alamos Neutron Science Center (LANSCE). The experimental method combines a prior experiment's characterization of prompt fission γ rays in conjunction with a fission tagging detector with a separate experiment using a thick 239Pu sample to extract the neutron capture cross section in ratio to 239Pu (n, f). We have made new predictions of the capture cross section taking into account recent results for the M1more » scissors mode present in other actinides. The results show deviations from current evaluations which are 30% higher at the highest <span class="hlt">energies</span>, and will be used to improve calculations relevant for several applications.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2005NIMPB.241..521P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2005NIMPB.241..521P"><span>Study of the surface activation of ETFE by low <span class="hlt">energy</span> (<span class="hlt">keV</span>) Si and N bombardment</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Parada, M. A.; de Almeida, A.; Muntele, C.; Muntele, I.; Delalez, N.; Ila, D.</p> <p>2005-12-01</p> <p>The ethylenetetrafluoroethylene (ETFE) is a polymer formed by alternating ethylene and tetrafluoroethylene segments. It can be applied in the field of medical physics as intra venous catheters and as radiation dosimeters. The increasing application of polymeric materials in technological and scientific fields has motivated the use of surface treatments to modify the physical and chemical properties of polymer surfaces. When a material is exposed to ionizing radiation, it suffers damage leading to surface activation depending on the type, <span class="hlt">energy</span> and intensity of the applied radiation. In order to determine the radiation damage and the surface activation mechanism ETFE films were bombarded with <span class="hlt">keV</span> Si and N at various fluences. The bombarded film was also analyzed with optical absorption photospectrometry (OAP), Raman and Fourier transform infrared-attenuated total reflectance (FTIR-ATR) spectroscopy showing quantitatively the chemical nature at the damage caused by the Si and N bombardment.</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> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22382001-where-do-kev-photons-come-from-morphological-study-galactic-center-perseus','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22382001-where-do-kev-photons-come-from-morphological-study-galactic-center-perseus"><span>Where do the 3.5 <span class="hlt">keV</span> photons come from? A morphological study of the Galactic Center and of Perseus</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>Carlson, Eric; Jeltema, Tesla; Profumo, Stefano, E-mail: erccarls@ucsc.edu, E-mail: tesla@ucsc.edu, E-mail: profumo@ucsc.edu</p> <p></p> <p>We test the origin of the 3.5 <span class="hlt">keV</span> line photons by analyzing the morphology of the emission at that <span class="hlt">energy</span> from the Galactic Center and from the Perseus cluster of galaxies. We employ a variety of different templates to model the continuum emission and analyze the resulting radial and azimuthal distribution of the residual emission. We then perform a pixel-by-pixel binned likelihood analysis including line emission templates and dark matter templates and assess the correlation of the 3.5 <span class="hlt">keV</span> emission with these templates. We conclude that the radial and azimuthal distribution of the residual emission is incompatible with a darkmore » matter origin for both the Galactic center and Perseus; the Galactic center 3.5 <span class="hlt">keV</span> line photons trace the morphology of lines at comparable <span class="hlt">energy</span>, while the Perseus 3.5 <span class="hlt">keV</span> photons are highly correlated with the cluster's cool core, and exhibit a morphology incompatible with dark matter decay. The template analysis additionally allows us to set the most stringent constraints to date on lines in the 3.5 <span class="hlt">keV</span> range from dark matter decay.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018MS%26E..298a2020A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018MS%26E..298a2020A"><span>Mass attenuation coefficients of several bio-adhesive based oil palm particleboards at 16.59-25.26 <span class="hlt">keV</span> photon <span class="hlt">energies</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Abdu Mustapa, U. A.; Yusof, M. F. Mohd; Hamid, P. N. K. Abd; Hashim, R.; Ahmad, M. Z.; Aziz, M. Z. Abd</p> <p>2018-01-01</p> <p>Particleboards made of oil palm with addition of polylactic acid (PLA), starch, and fish oil were fabricated with target density of 1.0 g/cm3. The mass attenuation coefficients of the particleboards were measured using x-ray fluorescence (XRF) configuration in conjunction with niobium, molybdenum, palladium and tin metal plates that provided Kα1 photon <span class="hlt">energies</span> between 16.59 and 25.26 <span class="hlt">keV</span>. The results were compared to the calculated value of water using XCOM. The results showed that all particleboards having mass attenuation coefficients near to the value of water with the mass attenuation coefficient different less than 0.25. The method of fabrication did not give significant different to the mass attenuation coefficients of the particleboards. The results had indicated the potential of bio-adhesive based palm oil particleboards to be developed as phantoms for low <span class="hlt">energy</span> photons.</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('http://adsabs.harvard.edu/abs/1998RaPC...52...73H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1998RaPC...52...73H"><span>Sterilization of foods with low-<span class="hlt">energy</span> electrons (``soft-electrons'')</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hayashi, Toru; Takahashi, Yoko; Todoriki, Setsuko</p> <p>1998-06-01</p> <p>Electrons with an <span class="hlt">energy</span> of 300 <span class="hlt">keV</span> or lower were defined as "Soft-electrons", which showed several advantages over conventional irradiation with gamma-rays or high-<span class="hlt">energy</span> electrons in decontamination of grains and spices. <span class="hlt">Energies</span> of electrons necessary to reduce microbial loads to levels lower than 10 CFU/g were 60 <span class="hlt">keV</span> for brown rice, 75 <span class="hlt">keV</span> for wheat, 100 <span class="hlt">keV</span> for white pepper, coriander and basil, 130 <span class="hlt">keV</span> for buckwheat, 160 <span class="hlt">keV</span> for rough rice, and 210 <span class="hlt">keV</span> for black pepper. Electrons with such <span class="hlt">energies</span> did not significantly influence the quality.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016PASJ...68S..30O','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016PASJ...68S..30O"><span>Suzaku Wide-<span class="hlt">band</span> All-sky Monitor measurements of duration distributions of gamma-ray bursts</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ohmori, Norisuke; Yamaoka, Kazutaka; Ohno, Masanori; Sugita, Satoshi; Kinoshita, Ryuuji; Nishioka, Yusuke; Hurley, Kevin; Hanabata, Yoshitaka; Tashiro, Makoto S.; Enomoto, Junichi; Fujinuma, Takeshi; Fukazawa, Yasushi; Iwakiri, Wataru; Kawano, Takafumi; Kokubun, Motohide; Makishima, Kazuo; Matsuoka, Shunsuke; Nagayoshi, Tsutomu; Nakagawa, Yujin E.; Nakaya, Souhei; Nakazawa, Kazuhiro; Takahashi, Tadayuki; Takeda, Sawako; Terada, Yukikatsu; Urata, Yuji; Yabe, Seiya; Yasuda, Tetsuya; Yamauchi, Makoto</p> <p>2016-06-01</p> <p>We report on the T90 and T50 duration distributions and their relations with spectral hardness using 1464 gamma-ray bursts (GRBs), which were observed by the Suzaku Wide-<span class="hlt">band</span> All-sky Monitor (WAM) from 2005 August 4 to 2010 December 29. The duration distribution is clearly bimodal in three <span class="hlt">energy</span> ranges (50-120, 120-250, and 250-550 <span class="hlt">keV</span>), but is unclear in the 550-5000 <span class="hlt">keV</span> range, probably because of the limited sample size. The WAM durations decrease with <span class="hlt">energy</span> according to a power-law index of -0.058(-0.034, +0.033). The hardness-duration relation reveals the presence of short-hard and long-soft GRBs. The short:long event ratio tends to be higher with increasing <span class="hlt">energy</span>. We compared the WAM distribution with ones measured by eight other GRB instruments. The WAM T90 distribution is very similar to those of INTEGRAL/SPI-ACS and Granat/PHEBUS, and least likely to match the Swift/BAT distribution. The WAM short:long event ratio (0.25:0.75) is much different from Swift/BAT (0.08:0.92), but is almost the same as CGRO/BATSE (0.25:0.75). To explain this difference for BAT, we examined three effects: BAT trigger types, <span class="hlt">energy</span> dependence of the duration, and detection sensitivity differences between BAT and WAM. As a result, we found that the ratio difference could be explained mainly by <span class="hlt">energy</span> dependence including soft extended emissions for short GRBs and much better sensitivity for BAT which can detect weak/long GRBs. The reason for the same short:long event ratio for BATSE and WAM was confirmed by calculation using the trigger efficiency curve.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22047895-spectral-properties-approx-kev-energetic-neutral-atoms-measured-interstellar-boundary-explorer-ibex-along-lines-sight-voyager','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22047895-spectral-properties-approx-kev-energetic-neutral-atoms-measured-interstellar-boundary-explorer-ibex-along-lines-sight-voyager"><span>SPECTRAL PROPERTIES OF {approx}0.5-6 <span class="hlt">keV</span> ENERGETIC NEUTRAL ATOMS MEASURED BY THE INTERSTELLAR BOUNDARY EXPLORER (IBEX) ALONG THE LINES OF SIGHT OF VOYAGER</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>Desai, M. I.; Allegrini, F. A.; Dayeh, M. A.</p> <p>2012-04-20</p> <p>Energetic neutral atoms (ENAs) observed by the Interstellar Boundary Explorer (IBEX) provide powerful diagnostics about the origin of the progenitor ion populations and the physical mechanisms responsible for their production. Here we survey the fluxes, <span class="hlt">energy</span> spectra, and <span class="hlt">energy</span> dependence of the spectral indices of {approx}0.5-6 <span class="hlt">keV</span> ENAs measured by IBEX-Hi along the lines of sight of Voyager 1 and 2. We compare the ENA spectra observed at IBEX with predictions of Zank et al. who modeled the microphysics of the heliospheric termination shock to predict the shape and relative contributions of three distinct heliosheath ion populations. We show thatmore » (1) the ENA spectral indices exhibit similar <span class="hlt">energy</span> dependence along V1 and V2 directions-the spectrum hardens to {gamma} {approx} 1 between {approx}1 and 2 <span class="hlt">keV</span> and softens to {gamma} {approx} 2 below {approx}1 <span class="hlt">keV</span> and above {approx}2 <span class="hlt">keV</span>, (2) the observed ENA fluxes agree to within {approx}50% of the Zank et al. predictions and are unlikely to be produced by core solar wind (SW) ions, and (3) the ENA spectra do not exhibit sharp cutoffs at {approx}twice the SW speed as is typically observed for shell-like pickup ion (PUI) distributions in the heliosphere. We conclude that ENAs at IBEX are generated by at least two types of ion populations whose relative contributions depend on the ENA <span class="hlt">energy</span>: transmitted PUIs in the {approx}0.5-5 <span class="hlt">keV</span> <span class="hlt">energy</span> range and reflected PUIs above {approx}5 <span class="hlt">keV</span> <span class="hlt">energy</span>. The {approx}0.5-5 <span class="hlt">keV</span> PUI distribution is probably a superposition of Maxwellian or kappa distributions and partially filled shell distributions in velocity space.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017RScI...88j3503S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017RScI...88j3503S"><span>A 7.2 <span class="hlt">keV</span> spherical x-ray crystal backlighter for two-frame, two-color backlighting at Sandia's Z Pulsed Power Facility</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Schollmeier, M. S.; Knapp, P. F.; Ampleford, D. J.; Harding, E. C.; Jennings, C. A.; Lamppa, D. C.; Loisel, G. P.; Martin, M. R.; Robertson, G. K.; Shores, J. E.; Smith, I. C.; Speas, C. S.; Weis, M. R.; Porter, J. L.; McBride, R. D.</p> <p>2017-10-01</p> <p>Many experiments on Sandia National Laboratories' Z Pulsed Power Facility—a 30 MA, 100 ns rise-time, pulsed-power driver—use a monochromatic quartz crystal backlighter system at 1.865 <span class="hlt">keV</span> (Si He α ) or 6.151 <span class="hlt">keV</span> (Mn He α ) x-ray <span class="hlt">energy</span> to radiograph an imploding liner (cylindrical tube) or wire array z-pinch. The x-ray source is generated by the Z-Beamlet laser, which provides two 527-nm, 1 kJ, 1-ns laser pulses. Radiographs of imploding, thick-walled beryllium liners at convergence ratios CR above 15 [ C R = r i ( 0 ) / r i ( t ) ] using the 6.151-<span class="hlt">keV</span> backlighter system were too opaque to identify the inner radius r i of the liner with high confidence, demonstrating the need for a higher-<span class="hlt">energy</span> x-ray radiography system. Here, we present a 7.242 <span class="hlt">keV</span> backlighter system using a Ge(335) spherical crystal with the Co He α resonance line. This system operates at a similar Bragg angle as the existing 1.865 <span class="hlt">keV</span> and 6.151 <span class="hlt">keV</span> backlighters, enhancing our capabilities for two-color, two-frame radiography without modifying the system integration at Z. The first data taken at Z include 6.2-<span class="hlt">keV</span> and 7.2-<span class="hlt">keV</span> two-color radiographs as well as radiographs of low-convergence (CR about 4-5), high-areal-density liner implosions.</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('http://adsabs.harvard.edu/abs/2008ICRC....1...19W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2008ICRC....1...19W"><span>The High <span class="hlt">Energy</span> Photons Emission from Solar Flares Observed by SZ2-XD</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wang, Huanyu; Li, Xinqiao; Ma, Yuqian; Zhang, Chengmo; Xu, Yupeng; Wang, Jingzhou; Chen, Guoming</p> <p></p> <p>The spectra and light curve of near a hundred Solar X-ray Flare events, which were observed by SZ2/XD in the <span class="hlt">energy</span> <span class="hlt">band</span> of 10-800 <span class="hlt">keV</span> during 2001, have been investigated. The events covered from C to X-class flares, which are shown different characters of high <span class="hlt">energy</span> photons emission. The results will be presented in this paper. The discussions will be made especially for 3 of the brightest X-class solar flares SF010402(X20),SF010406(X5.6) and SF010415 (X14.4, a GLE event).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018RuPhJ..60.2201Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018RuPhJ..60.2201Z"><span>A Combined Model of Charging of the Surface and Bulk of a Dielectric Target by Electrons with the <span class="hlt">Energies</span> 10-30 <span class="hlt">keV</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zykov, V. M.; Neiman, D. A.</p> <p>2018-04-01</p> <p>A physico-mathematical model of the processes of radiation-induced charging of dielectric materials with open surfaces, irradiated with monoenergetic electrons in the <span class="hlt">energy</span> range 10-30 <span class="hlt">keV</span>, is described. The model takes into account the relationship between the processes of surface and bulk charging for the given conditions of the experimental design, which accounts for the effect of anomalously long charging of dielectrics after the incident <span class="hlt">energy</span> of primary electrons during charging is reduced to below the second critical <span class="hlt">energy</span> for the secondary electronic emission coefficient. The initial fast phase of charging a high-resistivity dielectric material (Al2O3) is investigated. It is shown that as the incident electron <span class="hlt">energy</span> is approaching the second critical <span class="hlt">energy</span> during charging, the secondary electronic emission is partially suppressed due to negative charging of the open surface of the dielectric and formation of a near-surface inversion electrical field retarding the electronic emission yield.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20150023328&hterms=Nebula&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3DNebula','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20150023328&hterms=Nebula&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3DNebula"><span>High-<span class="hlt">Energy</span> X-Ray Imaging of the Pulsar Wind Nebula MSH 15-52: Constraints on Particle Acceleration and Transport</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>An, Hongjun; Madsen, Kristin K.; Reynolds, Stephen P.; Kaspi, Victoria M.; Harrison, Fiona A.; Boggs, Steven E.; Christensen, Finn E.; Craig, William W.; Fryer, Chris L.; Grefenstette, Brian W.; <a style="text-decoration: none; " href="javascript:void(0); " onClick="displayelement('author_20150023328'); toggleEditAbsImage('author_20150023328_show'); toggleEditAbsImage('author_20150023328_hide'); "> <img style="display:inline; width:12px; height:12px; " src="images/arrow-up.gif" width="12" height="12" border="0" alt="hide" id="author_20150023328_show"> <img style="width:12px; height:12px; display:none; " src="images/arrow-down.gif" width="12" height="12" border="0" alt="hide" id="author_20150023328_hide"></p> <p>2014-01-01</p> <p>We present the first images of the pulsar wind nebula (PWN) MSH 15-52 in the hard X-ray <span class="hlt">band</span> (8 <span class="hlt">keV</span>), as measured with the Nuclear Spectroscopic Telescope Array (NuSTAR). Overall, the morphology of the PWN as measured by NuSTAR in the 3-7 <span class="hlt">keV</span> <span class="hlt">band</span> is similar to that seen in Chandra high-resolution imaging. However, the spatial extent decreases with <span class="hlt">energy</span>, which we attribute to synchrotron <span class="hlt">energy</span> losses as the particles move away from the shock. The hard-<span class="hlt">band</span> maps show a relative deficit of counts in the northern region toward the RCW 89 thermal remnant, with significant asymmetry. We find that the integrated PWN spectra measured with NuSTAR and Chandra suggest that there is a spectral break at 6 <span class="hlt">keV</span>, which may be explained by a break in the synchrotron emitting electron distribution at approximately 200 TeV and/or imperfect cross calibration. We also measure spatially resolved spectra, showing that the spectrum of the PWN softens away from the central pulsar B1509-58, and that there exists a roughly sinusoidal variation of spectral hardness in the azimuthal direction. We discuss the results using particle flow models. We find non-monotonic structure in the variation with distance of spectral hardness within 50 of the pulsar moving in the jet direction, which may imply particle and magnetic-field compression by magnetic hoop stress as previously suggested for this source. We also present two-dimensional maps of spectral parameters and find an interesting shell-like structure in the N(sub H) map. We discuss possible origins of the shell-like structure and their implications.</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('http://adsabs.harvard.edu/abs/2018PhRvC..97f5802L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PhRvC..97f5802L"><span>Determination of 20Ne(p ,γ )21Na cross sections from Ep=500 -2000 <span class="hlt">keV</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lyons, S.; Görres, J.; deBoer, R. J.; Stech, E.; Chen, Y.; Gilardy, G.; Liu, Q.; Long, A. M.; Moran, M.; Robertson, D.; Seymour, C.; Vande Kolk, B.; Wiescher, M.; Best, A.</p> <p>2018-06-01</p> <p>Background: The reaction 20Ne(p ,γ )21Na influences the nucleosynthesis of Ne, Na, and Mg isotopes while contributing to hydrogen burning in several stellar sites, such as red giants, asymptotic giant branch (AGB) stars, massive stars, and oxygen-neon (ONe) novae. In the relevant temperature range for these environments (T = 0.05-0.5 GK), the main contributions to this reaction rate are from the direct capture process as well as the high-<span class="hlt">energy</span> tail of a subthreshold resonance in the ground-state transition at Ex = 2425 <span class="hlt">keV</span> in the 21Na compound nucleus. Purpose: The previous measurement of this reaction reports cross sections with large uncertainties for the ground-state transition. At higher <span class="hlt">energies</span>, where the subthreshold resonance makes a smaller contribution to the total cross section, only upper limits are provided. This work aims to reduce the uncertainty in the cross section where direct capture dominates, as well as provide cross-section data in previously unmeasured regions. Method: The 20Ne(p ,γ )21Na reaction was measured over a wide proton <span class="hlt">energy</span> range (Ep = 0.5-2.0 MeV) at θlab = 90∘. Transitions to the ground state and to the 332 and 2425 <span class="hlt">keV</span> excited states were observed. The primary transitions to these three bound states were utilized in an R -matrix analysis to determine the contributions of the direct capture and the subthreshold resonance to the total cross section. Results: The cross sections of the present measurements have been found to be in good agreement with the previous data at low <span class="hlt">energy</span>. Significantly improved cross-section measurements have been obtained over the Ep = 1300-1900 <span class="hlt">keV</span> region. The narrow resonance at Ec.m. = 1113 <span class="hlt">keV</span> (Ex = 3544.3 <span class="hlt">keV</span>) has also been remeasured and its strength has been found to be in good agreement with previous measurements. Conclusions: An extrapolation of the S factor of 20Ne(p ,γ )21Na has been made to low <span class="hlt">energies</span> using the R -matrix fit. The reaction rate from the subthreshold resonance was</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017ChA%26A..41...42W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017ChA%26A..41...42W"><span>Periodicity Analysis of X-ray Light Curves of SS 433</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wang, Jun-yi; Lu, Xiang-long; Zhao, Qiu-wen; Dong, Dian-qiao; Lao, Bao-qiang; Lu, Yang; Wei, Yan-heng; Wu, Xiao-cong; An, Tao</p> <p>2017-01-01</p> <p>SS 433 is sofar the unique X-ray binary that has the simultaneously detected orbital period, super-orbital period, and nutation period, as well as a bidirectional spiral jet. The study on its X-ray light variability is helpful for understanding the dynamic process of the system, and the correlations between the different wavebands. In this paper, two time-series analysis techniques, i.e., the Lomb-Scargle periodogram and weighted wavelet Z-transform, are employed to search for the periods in the Swift/BAT (Burst Alert Telescope) (15-50 <span class="hlt">keV</span>) and RXTE/ASM (Rose X-ray Timing Explorer/All Sky Monitor) (1.5-3, 3- 4, and 5-12 <span class="hlt">keV</span>) light curves of SS 433, and the Monte Carlo simulation is performed for the obtained periodical components. For the 15-50 <span class="hlt">keV</span> <span class="hlt">energy</span> <span class="hlt">band</span>, five significant periodical components are detected, which are P1(∼6.29 d), P2 (∼6.54 d), P3 (∼13.08 d), P4 (∼81.50 d), and P5 (∼162.30 d). For the 3-5 and 5-12 <span class="hlt">keV</span> <span class="hlt">energy</span> <span class="hlt">bands</span>, the periodical components P3 (∼13 d) and P5 (∼162 d) are detected in both <span class="hlt">energy</span> <span class="hlt">bands</span>. However, for the 1.5-3 <span class="hlt">keV</span> <span class="hlt">energy</span> <span class="hlt">band</span>, no significant periodic signal is detected. P5 is the strongest periodic signal in the power spectrum for all the <span class="hlt">energy</span> <span class="hlt">bands</span> of 3-5, 5-12, and 15-50 <span class="hlt">keV</span>, and it is consistent with the previous result obtained from the study of optical light curves. Furthermore, in combination with the radio spiral jet of SS 433, it is suggested that the X-ray and optical variability of P5 (∼162 d) is probably related to the precession of its relativistic jet. The high correlation between the X-ray and optical light curves may also imply that the X-ray and optical radiations are of the same physical origin. P3 shows a good agreement with the orbital period (∼13.07 d) obtained by the previous study, and P2 and P4 are respectively the high-frequency harmonics of P3 and P5. P1 is detected only in the power spectrum of the 15-50 <span class="hlt">keV</span> <span class="hlt">energy</span> <span class="hlt">band</span>, and it is consistent with the nutation period of the system. As</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('https://www.ncbi.nlm.nih.gov/pubmed/26082026','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26082026"><span>Investigation of the effective atomic numbers of dosimetric materials for electrons, protons and alpha particles using a direct method in the <span class="hlt">energy</span> region 10 <span class="hlt">keV</span>-1 GeV: a comparative study.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kurudirek, Murat; Aksakal, Oğuz; Akkuş, Tuba</p> <p>2015-11-01</p> <p>A direct method has been used for the first time, to compute effective atomic numbers (Z eff) of water, air, human tissues, and some organic and inorganic compounds, for total electron proton and alpha particle interaction in the <span class="hlt">energy</span> region 10 <span class="hlt">keV</span>-1 GeV. The obtained values for Z eff were then compared to those obtained using an interpolation procedure. In general, good agreement has been observed for electrons, and the difference (%) in Z eff between the results of the direct and the interpolation method was found to be <10 % for all materials, in the <span class="hlt">energy</span> range from 10 <span class="hlt">keV</span> to 1 MeV. More specifically, results of the two methods were found to agree well (Dif. <10 %) for air, calcium fluoride, kapton polyimide film, paraffin wax and plastic scintillator in the entire <span class="hlt">energy</span> region with respect to the total electron interaction. On the other hand, values for Z eff calculated using both methods for protons and alpha particles generally agree with each other in the high-<span class="hlt">energy</span> region above 10 MeV.</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('https://www.osti.gov/pages/biblio/1441469-kev-spherical-ray-crystal-backlighter-two-frame-two-color-backlighting-sandias-pulsed-power-facility','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1441469-kev-spherical-ray-crystal-backlighter-two-frame-two-color-backlighting-sandias-pulsed-power-facility"><span>A 7.2 <span class="hlt">keV</span> spherical x-ray crystal backlighter for two-frame, two-color backlighting at Sandia’s Z Pulsed Power Facility</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Schollmeier, M. S.; Knapp, P. F.; Ampleford, D. J.; ...</p> <p>2017-10-10</p> <p>Many experiments on Sandia National Laboratories’ Z Pulsed Power Facility—a 30 MA, 100 ns rise-time, pulsed-power driver—use a monochromatic quartz crystal backlighter system at 1.865 <span class="hlt">keV</span> (Si He α) or 6.151 <span class="hlt">keV</span> (Mn He α) x-ray <span class="hlt">energy</span> to radiograph an imploding liner (cylindrical tube) or wire array z-pinch. The x-ray source is generated by the Z-Beamlet laser, which provides two 527-nm, 1 kJ, 1-ns laser pulses. Radiographs of imploding, thick-walled beryllium liners at convergence ratios C R above 15 [C R=r i(0)/r i(t)] using the 6.151-<span class="hlt">keV</span> backlighter system were too opaque to identify the inner radius ri of the linermore » with high confidence, demonstrating the need for a higher-<span class="hlt">energy</span> x-ray radiography system. Here, we present a 7.242 <span class="hlt">keV</span> backlighter system using a Ge(335) spherical crystal with the Co He α resonance line. This system operates at a similar Bragg angle as the existing 1.865 <span class="hlt">keV</span> and 6.151 <span class="hlt">keV</span> backlighters, enhancing our capabilities for two-color, two-frame radiography without modifying the system integration at Z. The first data taken at Z include 6.2-<span class="hlt">keV</span> and 7.2-<span class="hlt">keV</span> two-color radiographs as well as radiographs of low-convergence (C R about 4-5), high-areal-density liner implosions.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1441469','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1441469"><span>A 7.2 <span class="hlt">keV</span> spherical x-ray crystal backlighter for two-frame, two-color backlighting at Sandia’s Z Pulsed Power Facility</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>Schollmeier, M. S.; Knapp, P. F.; Ampleford, D. J.</p> <p></p> <p>Many experiments on Sandia National Laboratories’ Z Pulsed Power Facility—a 30 MA, 100 ns rise-time, pulsed-power driver—use a monochromatic quartz crystal backlighter system at 1.865 <span class="hlt">keV</span> (Si He α) or 6.151 <span class="hlt">keV</span> (Mn He α) x-ray <span class="hlt">energy</span> to radiograph an imploding liner (cylindrical tube) or wire array z-pinch. The x-ray source is generated by the Z-Beamlet laser, which provides two 527-nm, 1 kJ, 1-ns laser pulses. Radiographs of imploding, thick-walled beryllium liners at convergence ratios C R above 15 [C R=r i(0)/r i(t)] using the 6.151-<span class="hlt">keV</span> backlighter system were too opaque to identify the inner radius ri of the linermore » with high confidence, demonstrating the need for a higher-<span class="hlt">energy</span> x-ray radiography system. Here, we present a 7.242 <span class="hlt">keV</span> backlighter system using a Ge(335) spherical crystal with the Co He α resonance line. This system operates at a similar Bragg angle as the existing 1.865 <span class="hlt">keV</span> and 6.151 <span class="hlt">keV</span> backlighters, enhancing our capabilities for two-color, two-frame radiography without modifying the system integration at Z. The first data taken at Z include 6.2-<span class="hlt">keV</span> and 7.2-<span class="hlt">keV</span> two-color radiographs as well as radiographs of low-convergence (C R about 4-5), high-areal-density liner implosions.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/21133997-calculations-stopping-powers-ev-kev-electrons-elemental-solids','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/21133997-calculations-stopping-powers-ev-kev-electrons-elemental-solids"><span>Calculations of stopping powers of 100 eV-30 <span class="hlt">keV</span> electrons in 31 elemental solids</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>Tanuma, S.; Powell, C. J.; Penn, D. R.</p> <p></p> <p>We present calculated electron stopping powers (SPs) for 31 elemental solids (Li, Be, glassy C, graphite, diamond, Na, Mg, K, Sc, Ti, V, Fe, Y, Zr, Nb, Mo, Ru, Rh, In, Sn, Cs, Gd, Tb, Dy, Hf, Ta, W, Re, Os, Ir, and Bi). These SPs were determined with an algorithm previously used for the calculation of electron inelastic mean free paths and from <span class="hlt">energy</span>-loss functions (ELFs) derived from experimental optical data. The SP calculations were made for electron <span class="hlt">energies</span> between 100 eV and 30 <span class="hlt">keV</span> and supplement our earlier SP calculations for ten additional solids (Al, Si, Cr, Ni,more » Cu, Ge, Pd, Ag, Pt, and Au). Plots of SP versus atomic number for the group of 41 solids show clear trends. Multiple peaks and shoulders are seen that result from the contributions of valence-electron and various inner-shell excitations. Satisfactory agreement was found between the calculated SPs and values from the relativistic Bethe SP equation with recommended values of the mean excitation <span class="hlt">energy</span> (MEE) for <span class="hlt">energies</span> above 10 <span class="hlt">keV</span>. We determined effective MEEs versus maximum excitation <span class="hlt">energy</span> from the ELFs for each solid. Plots of effective MEE versus atomic number showed the relative contributions of valence-electron and different core-electron excitations to the MEE. For a maximum excitation <span class="hlt">energy</span> of 30 <span class="hlt">keV</span>, our effective MEEs agreed well for Be, graphite, Na, Al, and Si with recommended MEEs; a difference for Li was attributed to sample oxidation in the SP measurements for the recommended MEE. Substantially different effective MEEs were found for the three carbon allotropes (graphite, diamond, and glassy C)« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012cosp...39.1138M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012cosp...39.1138M"><span>Ion distributions in RC at different <span class="hlt">energy</span> levels retrieved from TWINS ENA images by voxel CT tech</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ma, S. Y.; McComas, David; Xu, Liang; Goldstein, Jerry; Yan, Wei-Nan</p> <p>2012-07-01</p> <p>Distributions of energetic ions in the RC regions in different <span class="hlt">energy</span> levels are retrieved by using 3-D voxel CT inversion method from ENA measurements onboard TWINS constellation during the main phase of a moderate geomagnetic storm. It is assumed that the ion flux distribution in the RC is anisotropic in regard to pitch angle which complies with the adiabatic invariance of the magnetic moment as ion moving in the dipole magnetic mirror field. A semi-empirical model of the RC ion distribution in the magnetic equator is quoted to form the ion flux distribution shape at off-equatorial latitudes by mapping. For the concerned time interval, the two satellites of the TWINS flying in double Molnia orbits were located in nearly the same meridian plane at vantage points widely separated in magnetic local time, and both more than 5 RE geocentric distance from the Earth. The ENA data used in this study are differential fluxes averaged over 12 sweeps (corresponding to an interval of 16 min.) at different <span class="hlt">energy</span> levels ranging from about 1 to 100 <span class="hlt">keV</span>. The retrieved ion distributions show that in total the main part of the RC is located in the region with L value larger than 4, tending to increase at larger L. It reveals that there are two distinct dominant <span class="hlt">energy</span> <span class="hlt">bands</span> at which the ion fluxes are significantly larger magnitude than at other <span class="hlt">energy</span> levels, one is at lower level around 2 <span class="hlt">keV</span> and the other at higher level of 30-100 <span class="hlt">keV</span>. Furthermore, it is very interesting that the peak fluxes of the RC ions at the two <span class="hlt">energy</span> <span class="hlt">bands</span> occurred in different magnetic local time, low <span class="hlt">energy</span> ions appear preferentially in after midnight, while the higher <span class="hlt">energy</span> ions mainly distributed around midnight and pre-midnight. This new profile is worthy of further study and needs to be demonstrated by more cases.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016ApJ...824L...4K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016ApJ...824L...4K"><span>NuSTAR Resolves the First Dual AGN above 10 <span class="hlt">keV</span> in SWIFT J2028.5+2543</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Koss, Michael J.; Glidden, Ana; Baloković, Mislav; Stern, Daniel; Lamperti, Isabella; Assef, Roberto; Bauer, Franz; Ballantyne, David; Boggs, Steven E.; Craig, William W.; Farrah, Duncan; Fürst, Felix; Gandhi, Poshak; Gehrels, Neil; Hailey, Charles J.; Harrison, Fiona A.; Markwardt, Craig; Masini, Alberto; Ricci, Claudio; Treister, Ezequiel; Walton, Dominic J.; Zhang, William W.</p> <p>2016-06-01</p> <p>We have discovered heavy obscuration in the dual active galactic nucleus (AGN) in the Swift/Burst Alert Telescope (BAT) source SWIFT J2028.5+2543 using Nuclear Spectroscopic Telescope Array (NuSTAR). While an early XMM-Newton study suggested the emission was mainly from NGC 6921, the superior spatial resolution of NuSTAR above 10 <span class="hlt">keV</span> resolves the Swift/BAT emission into two sources associated with the nearby galaxies MCG +04-48-002 and NGC 6921 (z = 0.014) with a projected separation of 25.3 kpc (91″). NuSTAR's sensitivity above 10 <span class="hlt">keV</span> finds both are heavily obscured to Compton-thick levels (N H ≈ (1-2) × 1024 cm-2) and contribute equally to the BAT detection ({L}10-50 {keV}{{int}} ≈ 6 × 1042 erg s-1). The observed luminosity of both sources is severely diminished in the 2-10 <span class="hlt">keV</span> <span class="hlt">band</span> ({L} 2-10 {keV}{{obs}}\\lt 0.1× {L} 2-10 {keV}{{int}}), illustrating the importance of >10 <span class="hlt">keV</span> surveys like those with NuSTAR and Swift/BAT. Compared to archival X-ray data, MCG +04-48-002 shows significant variability (>3) between observations. Despite being bright X-ray AGNs, they are difficult to detect using optical emission-line diagnostics because MCG +04-48-002 is identified as a starburst/composite because of the high rates of star formation from a luminous infrared galaxy while NGC 6921 is only classified as a LINER using line detection limits. SWIFT J2028.5+2543 is the first dual AGN resolved above 10 <span class="hlt">keV</span> and is the second most heavily obscured dual AGN discovered to date in the X-rays other than NGC 6240.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22375759-decaying-vector-dark-matter-explanation-kev-line-from-galaxy-clusters','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22375759-decaying-vector-dark-matter-explanation-kev-line-from-galaxy-clusters"><span>Decaying vector dark matter as an explanation for the 3.5 <span class="hlt">keV</span> line from galaxy clusters</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>Farzan, Yasaman; Akbarieh, Amin Rezaei, E-mail: yasaman@theory.ipm.ac.ir, E-mail: am_rezaei@physics.sharif.ir</p> <p>2014-11-01</p> <p>We present a Vector Dark Matter (VDM) model that explains the 3.5 <span class="hlt">keV</span> line recently observed in the XMM-Newton observatory data from galaxy clusters. In this model, dark matter is composed of two vector bosons, V and V', which couple to the photon through an effective generalized Chern-Simons coupling, g{sub V}. V' is slightly heavier than V with a mass splitting m{sub V'} – m{sub V} ≅ 3.5 <span class="hlt">keV</span>. The decay of V' to V and a photon gives rise to the 3.5 <span class="hlt">keV</span> line. The production of V and V' takes place in the early universe within the freeze-in framework through the effectivemore » g{sub V} coupling when m{sub V'} < T < Λ, Λ being the cut-off above which the effective g{sub V} coupling is not valid. We introduce a high <span class="hlt">energy</span> model that gives rise to the g{sub V} coupling at low <span class="hlt">energies</span>. To do this, V and V' are promoted to gauge bosons of spontaneously broken new U(1){sub V} and U(1){sub V'} gauge symmetries, respectively. The high <span class="hlt">energy</span> sector includes milli-charged chiral fermions that lead to the g{sub V} coupling at low <span class="hlt">energy</span> via triangle diagrams.« less</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('https://ntrs.nasa.gov/search.jsp?R=19720048295&hterms=pitch+detector&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Dpitch%2Bdetector','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19720048295&hterms=pitch+detector&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Dpitch%2Bdetector"><span>Anisotropic pitch angle distribution of 50 eV to 50 <span class="hlt">keV</span> particles at synchronous altitude.</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Deforest, S. E.; Mcilwain, C. F.</p> <p>1972-01-01</p> <p>At times, the electron pitch angle distributions at synchronous orbit have been observed to be highly anisotropic. In the local morning region, distributions concentrated near 90 deg are often observed in particles of less than approximately 2000 V. This anisotropy decreases with increasing <span class="hlt">energy</span> from 1 <span class="hlt">keV</span> to the detector's limit at 50 <span class="hlt">keV</span>. The time development of anisotropy is consistent with production by pitch angle scattering processes which are not effective on electrons with small velocities parallel to the magnetic field. Another type of distribution has been observed with the low-<span class="hlt">energy</span> (below 1000 V) electrons concentrated parallel and antiparallel to the magnetic field. These distributions are only seen in the dusk sector, but this may be an orbital artifact.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1165906-kev-ray-spectrum-circinus-galaxy-nustar-xmm-newton-chandra-fully-compton-thick-active-galactic-nucleus','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1165906-kev-ray-spectrum-circinus-galaxy-nustar-xmm-newton-chandra-fully-compton-thick-active-galactic-nucleus"><span>The 2-79 <span class="hlt">keV</span> X-ray Spectrum of the Circinus Galaxy with NuSTAR, XMM-Newton and Chandra: a Fully Compton-Thick Active Galactic Nucleus</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Arevalo, P.; Bauer, F. E.; Puccetti, S.; ...</p> <p>2014-07-30</p> <p>Here, the Circinus galaxy is one of the closest obscured active galactic nuclei (AGNs), making it an ideal target for detailed study. Combining archival Chandra and XMM-Newton data with new NuSTAR observations, we model the 2-79 <span class="hlt">keV</span> spectrum to constrain the primary AGN continuum and to derive physical parameters for the obscuring material. Chandra's high angular resolution allows a separation of nuclear and off-nuclear galactic emission. In the off-nuclear diffuse emission, we find signatures of strong cold reflection, including high equivalent-width neutral Fe lines. This Compton-scattered off-nuclear emission amounts to 18% of the nuclear flux in the Fe line region,more » but becomes comparable to the nuclear emission above 30 <span class="hlt">keV</span>. The new analysis no longer supports a prominent transmitted AGN component in the observed <span class="hlt">band</span>. We find that the nuclear spectrum is consistent with Compton scattering by an optically thick torus, where the intrinsic spectrum is a power law of photon index Γ = 2.2-2.4, the torus has an equatorial column density of N H = (6-10) × 10 24 cm –2, and the intrinsic AGN 2-10 <span class="hlt">keV</span> luminosity is (2.3-5.1) × 10 42 erg s –1. These values place Circinus along the same relations as unobscured AGNs in accretion rate versus Γ and L X versus L IR phase space. NuSTAR's high sensitivity and low background allow us to study the short timescale variability of Circinus at X-ray <span class="hlt">energies</span> above 10 <span class="hlt">keV</span> for the first time. Here, the lack of detected variability favors a Compton-thick absorber, in line with the spectral fitting results.« less</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://hdl.handle.net/2060/20080003788','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20080003788"><span>The Morphology of the X-ray Emission above 2 <span class="hlt">keV</span> from Jupiter's Aurorae</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Elsner, R.; Branduardi-Raymont, G.; Galand, M.; Grodent, D.; Gladstone, G. R.; Waite, J. H.; Cravens, T.; Ford, P.</p> <p>2007-01-01</p> <p>The discovery in XMM-Newton X-ray data of X-ray emission above 2 keY from Jupiter's aurorae has led us to reexamine the Chandra ACIS-S observations taken in Feb 2003. Chandra's superior spatial resolution has revealed that the auroral X-rays with E > 2 <span class="hlt">keV</span> are emitted from the periphery of the region emitting those with E < 1 <span class="hlt">keV</span>. We are presently exploring the relationship of this morphology to that of the FUV emission from the main auroral oval and the polar cap. The low <span class="hlt">energy</span> emission has previously been established as due to charge exchange between energetic precipitating ions of oxygen and either sulfur or carbon. It seems likely to us that the higher <span class="hlt">energy</span> emission is due to precipitation of energetic electrons, possibly the same population of electrons responsible for the FUV emission. We discuss our analysis and interpretation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1998APS..DNP..B601C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1998APS..DNP..B601C"><span>Test of 4-body Theory via Polarized p-T Capture Below 80 <span class="hlt">keV</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Canon, R. S.; Gaff, S. J.; Kelley, J. H.; Schreiber, E. C.; Weller, H. R.; Wulf, E. A.; Prior, R. M.; Spraker, M.; Tilley, D. R.</p> <p>1998-10-01</p> <p>Our previous study of polarized p-d capture at <span class="hlt">energies</span> below 80 <span class="hlt">keV</span> revealed the major role played by MEC effects and provided a clean testing ground for state-of-the-art 3-body theory (the ``Ay puzzle'' remains)(G. Schmid et al); PRL 76, 3088(1996); PRC 56, 2565(1997). Four-body theory is on the threshold(A. Fonseca,W. Glöckle,A. Kievsky,H. Witala;Private communication) of being able to make similar ab-initio predictions. The p-T capture reaction is expected to exhibit strong MEC effects at very low <span class="hlt">energies</span> for reasons similar to those in p-d capture. Preliminary results indicate finite values of A_y(90^circ) in the 50-80 <span class="hlt">keV</span> region. These results will be discussed with respect to their implications on the M1 strength present in this reaction. Plans for future measurements and analysis will also be described.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013APS..DPPPO7006F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013APS..DPPPO7006F"><span>Demonstration of a 13 <span class="hlt">keV</span> Kr K-shell X-Ray Source at the National Ignition Facility</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Fournier, K. B.; May, M. J.; Colvin, J. D.; Barrios, M. A.; Patterson, J. R.; Regan, S. P.</p> <p>2013-10-01</p> <p>We report 3% conversion efficiency of laser <span class="hlt">energy</span> into Kr K-shell (~13 <span class="hlt">keV</span>) radiation, consistent with theoretical predictions. This is ~10 × greater than previous work. The emission was produced from a 4.1 mm diameter, 4 mm tall gas pipe target filled with 1.2 or 1.5 atm of Kr gas. 160 of the NIF laser beams deposited ~700 kJ of 3 ω light into the target in a ~140 TW, 5.0 ns duration square pulse. This laser configuration sufficiently heated the targets to optimize the K-shell x-ray emission. The Dante diagnostics measured ~5 TW into 4 π solid angle of >=12 <span class="hlt">keV</span> x rays for ~4 ns, which includes both continuum emission and flux in the Kr Heα line at 13 <span class="hlt">keV</span>. This work was performed under the auspices of the U.S. Department of <span class="hlt">Energy</span> by Lawrence Livermore National Laboratory under Contract DE-AC52-07NA27344. This work was supported by the Defense Threat Reduction Agency under the intera- gency agreements 10027-1420 and 10027-6167.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017EGUGA..19.2191W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EGUGA..19.2191W"><span>Quiet-Time Suprathermal (˜0.1 - 200 <span class="hlt">keV</span>) Electrons in the Solar Wind</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wang, Linghua; Yang, Liu; Tao, Jiawei; Zong, Qiugang; Li, Gang; Wimmer-Schweingruber, Robert; He, Jiansen; Tu, Chuanyi; Bale, Stuart</p> <p>2017-04-01</p> <p>We present a statistical survey of the <span class="hlt">energy</span> spectrum of solar wind suprathermal (˜0.1-200 <span class="hlt">keV</span>) electrons measured by the WIND 3DP instrument at 1 AU during quiet times at the minimum and maximum of solar cycles 23 and 24. The observed <span class="hlt">energy</span> spectrum of both (beaming) strahl and (isotropic) halo electrons at ˜0.1-1.5 <span class="hlt">keV</span> generally fits to a Kappa distribution function with an index κ and effective temperature Teff, while the observed <span class="hlt">energy</span> spectrum of nearly isotropic superhalo electrons at ˜20-200 <span class="hlt">keV</span> generally fits to a power-law function, J ˜ E-β. We find a strong positive correlation between κ and Teff for both strahl and halo electrons, and a strong positive correlation between the strahl density and halo density. In both solar cycles, κ is larger at solar minimum than at solar maximum for both strahl and halo electrons. For the superhalo population, the spectral index β ranges from ˜1.6 to ˜3.7 and the integrated density nsup ranges from 10-8 cm-3 to 10-5 cm-3, with no clear association with the sunspot number. In solar cycle 23 (24), the distribution of β has a broad maximum between 2.4 and 2.8 (2.0 and 2.4). All the strahl, halo and superhalo populations show no obvious correlation with the solar wind core population. These results reflect the nature of the generation of solar wind suprathermal electrons.</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('http://adsabs.harvard.edu/abs/2016AGUFMSH13D..05W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFMSH13D..05W"><span>Quiet-Time Suprathermal ( 0.1-1.5 <span class="hlt">keV</span>) Electrons in the Solar Wind</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wang, L.; Tao, J.; Zong, Q.; Li, G.; Salem, C. S.; Wimmer-Schweingruber, R. F.; He, J.; Tu, C.; Bale, S. D.</p> <p>2016-12-01</p> <p>We present a statistical survey of the <span class="hlt">energy</span> spectrum of solar wind suprathermal (˜0.1-1.5 <span class="hlt">keV</span>) electrons measured by the WIND/3DP instrument at 1 AU during quiet times at the minimum and maximum of solar cycles 23 and 24. After separating (beaming) strahl electrons from (isotropic) halo electrons according to their different behaviors in the angular distribution, we fit the observed <span class="hlt">energy</span> spectrum of both strahl and halo electrons at ˜0.1-1.5 <span class="hlt">keV</span> to a Kappa distribution function with an index κ and effective temperature Teff. We also calculate the number density n and average <span class="hlt">energy</span> Eavg of strahl and halo electrons by integrating the electron measurements between ˜0.1 and 1.5 <span class="hlt">keV</span>. We find a strong positive correlation between κ and Teff for both strahl and halo electrons, and a strong positive correlation between the strahl n and halo n, likely reflecting the nature of the generation of these suprathermal electrons. In both solar cycles, κ is larger at solar minimum than at solar maximum for both strahl and halo electrons. The halo κ is generally smaller than the strahl κ (except during the solar minimum of cycle 23). The strahl n is larger at solar maximum, but the halo n shows no difference between solar minimum and maximum. Both the strahl n and halo n have no clear association with the solar wind core population, but the density ratio between the strahl and halo roughly anti-correlates (correlates) with the solar wind density (velocity).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016ApJ...820...22T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016ApJ...820...22T"><span>Quiet-time Suprathermal (~0.1-1.5 <span class="hlt">keV</span>) Electrons in the Solar Wind</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Tao, Jiawei; Wang, Linghua; Zong, Qiugang; Li, Gang; Salem, Chadi S.; Wimmer-Schweingruber, Robert F.; He, Jiansen; Tu, Chuanyi; Bale, Stuart D.</p> <p>2016-03-01</p> <p>We present a statistical survey of the <span class="hlt">energy</span> spectrum of solar wind suprathermal (˜0.1-1.5 <span class="hlt">keV</span>) electrons measured by the WIND 3DP instrument at 1 AU during quiet times at the minimum and maximum of solar cycles 23 and 24. After separating (beaming) strahl electrons from (isotropic) halo electrons according to their different behaviors in the angular distribution, we fit the observed <span class="hlt">energy</span> spectrum of both strahl and halo electrons at ˜0.1-1.5 <span class="hlt">keV</span> to a Kappa distribution function with an index κ and effective temperature Teff. We also calculate the number density n and average <span class="hlt">energy</span> Eavg of strahl and halo electrons by integrating the electron measurements between ˜0.1 and 1.5 <span class="hlt">keV</span>. We find a strong positive correlation between κ and Teff for both strahl and halo electrons, and a strong positive correlation between the strahl n and halo n, likely reflecting the nature of the generation of these suprathermal electrons. In both solar cycles, κ is larger at solar minimum than at solar maximum for both strahl and halo electrons. The halo κ is generally smaller than the strahl κ (except during the solar minimum of cycle 23). The strahl n is larger at solar maximum, but the halo n shows no difference between solar minimum and maximum. Both the strahl n and halo n have no clear association with the solar wind core population, but the density ratio between the strahl and halo roughly anti-correlates (correlates) with the solar wind density (velocity).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19720018073','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19720018073"><span>A study of 2-20 <span class="hlt">KeV</span> X-rays from the Cygnus region</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Bleach, R. D.</p> <p>1972-01-01</p> <p>Two rocket-borne proportional counters, each with 650 sq c, met area and 1.8 x 7.1 deg FWHM rectangular mechanical collimation, surveyed the Cygnus region in the 2 to 20 <span class="hlt">keV</span> <span class="hlt">energy</span> range on two occasions. X-ray spectral data gathered on 21 September 1970 from discrete sources in Cygnus are presented. The data from Cyg X-1, Cyg X-2, and Cyg X-3 have sufficient statistical significance to indicate mutually exclusive spectral forms for the three. Upper limits are presented for X-ray intensities above 2 <span class="hlt">keV</span> for Cyg X-4 and Cyg X-5 (Cygnus loop). A search was made on 9 August 1971 for a diffuse component of X-rays 1.5 <span class="hlt">keV</span> associated with an interarm region of the galaxy at galactic longitudes in the vicinity of 60 degrees. A statistically significant excess associated with a narrow disk component was detected. Several possible emission models are discussed, with the most likely candidate being a population of unresolvable low luminosity discrete sources.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010EGUGA..12.8909M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010EGUGA..12.8909M"><span>Survey of upper <span class="hlt">band</span> chorus and ECH waves: Implications for the diffuse aurora</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Meredith, Nigel; Horne, Richard; Thorne, Richard; Anderson, Roger</p> <p>2010-05-01</p> <p>The origin of the diffuse aurora has been a source of controversy for many years. More recently the question has taken a new significance in view of the associated changes in atmospheric chemistry which may affect the middle atmosphere. Here we use CRRES data to assess the importance of upper <span class="hlt">band</span> chorus and electron cyclotron harmonic (ECH) waves in the production of the diffuse aurora. Both wave modes increase with increasing geomagnetic activity, suggesting they are related to periods of enhanced convection and/or substorm activity. They are confined to the near-equatorial region which excludes the pre-noon sector from the wave survey. During active conditions intense ECH waves and upper <span class="hlt">band</span> chorus, with amplitudes exceeding 1 mVm-1, are observed in the region 4 < L < 7 from 2100 to 0600 MLT approximately 20% and 6% of the time respectively. This suggests that both wave modes can put electrons on strong diffusion, but only during active conditions and not at all local times. Scattering rates fall below the strong diffusion limit at other times when the wave amplitudes are weaker. Fluxes of low <span class="hlt">energy</span> electrons (100 eV < E < 30 <span class="hlt">keV</span>) also increase with increasing geomagnetic activity in approximately the same region of geospace as the waves, suggesting that these electrons are responsible for the generation of the waves. The patterns of the upper <span class="hlt">band</span> chorus, ECH waves and low <span class="hlt">energy</span> electrons are similar to the global morphology of the diffuse aurora, suggesting that both wave modes play significant roles in the production of the diffuse aurora.</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/2017HEAD...1610927G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017HEAD...1610927G"><span>NuSTAR Hard X-ray Observations of the Energetic Millisecond Pulsars PSR B1821-24, PSR B1937+21, and PSR J0218+4232</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gotthelf, Eric V.; Bogdanov, Slavko</p> <p>2017-08-01</p> <p>We present NuSTAR hard X-ray timing and spectroscopy of the three exceptionally energetic rotation-powered millisecond pulsars PSRs B1821-24, B1937+21, and J0218+4232. By correcting for frequency and phase drifts of the NuSTAR on-board clock we are able to recover the intrinsic hard X-ray pulse profiles of all three pulsars with a resolution down to <15 ms. The substantial reduction of background emission relative to previous broad-<span class="hlt">band</span> X-ray observations allows us to detect for the first time pulsed emission up to ~50 <span class="hlt">keV</span>, ~20 <span class="hlt">keV</span>, and ~25 <span class="hlt">keV</span>, for the three pulsars, respectively. We conduct phase-resolved spectroscopy in the 0.5 - 79 <span class="hlt">keV</span> range for all three objects, obtaining the best yet measurements of the broad-<span class="hlt">band</span> spectral shape and high-<span class="hlt">energy</span> pulsed emission to date. We find extensions of the same power-law continua seen at lower <span class="hlt">energies</span>, with no conclusive evidence for a spectral turnover or break. Extrapolation of the X-ray power-law spectrum to higher <span class="hlt">energies</span> reveals that a turnover in the 100 <span class="hlt">keV</span> to 100 MeV range is required to accommodate the high <span class="hlt">energy</span> gamma-ray emission observed with Fermi LAT, similar to the broad-<span class="hlt">band</span> spectral <span class="hlt">energy</span> distribution observed for the Crab pulsar.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1997RScI...68..324C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1997RScI...68..324C"><span>Development and characterization of semiconductor ion detectors for plasma diagnostics in the range over 0.3 <span class="hlt">keV</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Cho, T.; Sakamoto, Y.; Hirata, M.; Kohagura, J.; Makino, K.; Kanke, S.; Takahashi, K.; Okamura, T.; Nakashima, Y.; Yatsu, K.; Tamano, T.; Miyoshi, S.</p> <p>1997-01-01</p> <p>For the purpose of plasma-ion-<span class="hlt">energy</span> analyses in a wide-<span class="hlt">energy</span> range from a few hundred eV to hundreds of <span class="hlt">keV</span>, upgraded semiconductor detectors are newly fabricated and characterized using a test-ion-beam line from 0.3 to 12 <span class="hlt">keV</span>. In particular, the detectable lowest-ion <span class="hlt">energy</span> is drastically improved at least down to 0.3 <span class="hlt">keV</span>; this <span class="hlt">energy</span> is one to two orders-of-magnitude better than those for commercially available Si-surface-barrier diodes employed for previous plasma-ion diagnostics. A signal-to-noise ratio of two to three orders-of-magnitude better than that for usual metal-collector detectors is demonstrated for the compact-sized semiconductor along with the availability of the use under conditions of a good vacuum and a strong-magnetic field. Such characteristics are achieved due to the improving methods of the optimization of the thicknesses of a Si dead layer and a SiO2 layer, as well as the nitrogen-doping technique near the depletion layer along with minimizing impurity concentrations in Si. Such an upgraded capability of an extremely low-<span class="hlt">energy</span>-ion detection with the low-noise characteristics enlarges research regimes of plasma-ion behavior using semiconductor detectors not only in the divertor regions of tokamaks but in wider spectra of open-field plasma devices including tandem mirrors. An application of the semiconductor ion detector for plasma-ion diagnostics is demonstrated in a specially designed ion-spectrometer structure.</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/2017PhPl...24j2708L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017PhPl...24j2708L"><span>Confined ion <span class="hlt">energy</span> >200 <span class="hlt">keV</span> and increased fusion yield in a DPF with monolithic tungsten electrodes and pre-ionization</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lerner, Eric J.; Hassan, Syed M.; Karamitsos, Ivana; Von Roessel, Fred</p> <p>2017-10-01</p> <p>To reduce impurities in the dense plasma focus FF-1 device, we used monolithic tungsten electrodes with pre-ionization. With this new set-up, we demonstrated a three-fold reduction of impurities by mass and a ten-fold reduction by ion number. FF-1 produced a 50% increase in fusion yield over our previous copper electrodes, both for a single shot and for a mean of ten consecutive shots with the same conditions. These results represent a doubling of fusion yield as compared with any other plasma focus device with the same 60 kJ <span class="hlt">energy</span> input. In addition, FF-1 produced a new single-shot record of 240 ± 20 <span class="hlt">keV</span> for mean ion <span class="hlt">energy</span>, a record for any confined fusion plasma, using any device, and a 50% improvement in ten-shot mean ion <span class="hlt">energy</span>. With a deuterium-nitrogen mix and corona-discharge pre-ionization, we were also able to reduce the standard deviation in the fusion yield to about 15%, a four-fold reduction over the copper-electrode results. We intend to further reduce impurities with new experiments using microwave treatment of tungsten electrodes, followed by the use of beryllium electrodes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014IJMPS..2760147R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014IJMPS..2760147R"><span>Ion Beam Materials Analysis and Modifications at <span class="hlt">keV</span> to MeV <span class="hlt">Energies</span> at the University of North Texas</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Rout, Bibhudutta; Dhoubhadel, Mangal S.; Poudel, Prakash R.; Kummari, Venkata C.; Lakshantha, Wickramaarachchige J.; Manuel, Jack E.; Bohara, Gyanendra; Szilasi, Szabolcs Z.; Glass, Gary A.; McDaniel, Floyd D.</p> <p>2014-02-01</p> <p>The University of North Texas (UNT) Ion Beam Modification and Analysis Laboratory (IBMAL) has four particle accelerators including a National Electrostatics Corporation (NEC) 9SDH-2 3 MV tandem Pelletron, a NEC 9SH 3 MV single-ended Pelletron, and a 200 kV Cockcroft-Walton. A fourth HVEC AK 2.5 MV Van de Graaff accelerator is presently being refurbished as an educational training facility. These accelerators can produce and accelerate almost any ion in the periodic table at <span class="hlt">energies</span> from a few <span class="hlt">keV</span> to tens of MeV. They are used to modify materials by ion implantation and to analyze materials by numerous atomic and nuclear physics techniques. The NEC 9SH accelerator was recently installed in the IBMAL and subsequently upgraded with the addition of a capacitive-liner and terminal potential stabilization system to reduce ion <span class="hlt">energy</span> spread and therefore improve spatial resolution of the probing ion beam to hundreds of nanometers. Research involves materials modification and synthesis by ion implantation for photonic, electronic, and magnetic applications, micro-fabrication by high <span class="hlt">energy</span> (MeV) ion beam lithography, microanalysis of biomedical and semiconductor materials, development of highenergy ion nanoprobe focusing systems, and educational and outreach activities. An overview of the IBMAL facilities and some of the current research projects are discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017RScI...88j3107S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017RScI...88j3107S"><span>X-ray spectrometer having 12 000 resolving power at 8 <span class="hlt">keV</span> <span class="hlt">energy</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Seely, John F.; Hudson, Lawrence T.; Henins, Albert; Feldman, Uri</p> <p>2017-10-01</p> <p>An x-ray spectrometer employing a thin (50 μm) silicon transmission crystal was used to record high-resolution Cu Kα spectra from a laboratory x-ray source. The diffraction was from the (331) planes that were at an angle of 13.26° to the crystal surface. The components of the spectral lines resulting from single-vacancy (1s) and double-vacancy (1s and 3d) transitions were observed. After accounting for the natural lifetime widths from reference double-crystal spectra and the spatial resolution of the image plate detector, the intrinsic broadening of the transmission crystal was measured to be as small as 0.67 eV and the resolving power 12 000, the highest resolving power achieved by a compact (0.5 m long) spectrometer employing a single transmission crystal operating in the hard x-ray region. By recording spectra with variable source-to-crystal distances and comparing to the calculated widths from various geometrical broadening mechanisms, the primary contributions to the intrinsic crystal broadening were found to be the source height at small distances and the crystal apertured height at large distances. By reducing these two effects, using a smaller source size and vignetting the crystal height, the intrinsic crystal broadening is then limited by the crystal thickness and the rocking curve width and would be 0.4 eV at 8 <span class="hlt">keV</span> <span class="hlt">energy</span> (20 000 resolving power).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016APS..DPPP10100T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016APS..DPPP10100T"><span>Spectral and Atomic Physics Analysis of Xenon L-Shell Emission From High <span class="hlt">Energy</span> Laser Produced Plasmas</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Thorn, Daniel; Kemp, G. E.; Widmann, K.; Benjamin, R. D.; May, M. J.; Colvin, J. D.; Barrios, M. A.; Fournier, K. B.; Liedahl, D.; Moore, A. S.; Blue, B. E.</p> <p>2016-10-01</p> <p>The spectrum of the L-shell (n =2) radiation in mid to high-Z ions is useful for probing plasma conditions in the multi-keV temperature range. Xenon in particular with its L-shell radiation centered around 4.5 <span class="hlt">keV</span> is copiously produced from plasmas with electron temperatures in the 5-10 <span class="hlt">keV</span> range. We report on a series of time-resolved L-shell Xe spectra measured with the NIF X-ray Spectrometer (NXS) in high-<span class="hlt">energy</span> long-pulse (>10 ns) laser produced plasmas at the National Ignition Facility. The resolving power of the NXS is sufficiently high (E/ ∂E >100) in the 4-5 <span class="hlt">keV</span> spectral <span class="hlt">band</span> that the emission from different charge states is observed. An analysis of the time resolved L-shell spectrum of Xe is presented along with spectral modeling by detailed radiation transport and atomic physics from the SCRAM code and comparison with predictions from HYDRA a radiation-hydrodynamics code with inline atomic-physics from CRETIN. This work was performed under the auspices of the U.S. Department of <span class="hlt">Energy</span> by LLNL under Contract DE-AC52-07NA27344.</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/2016ApJ...823..126A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016ApJ...823..126A"><span>Thermal Characteristics and the Differential Emission Measure Distribution During a B8.3 Flare on 2009 July 4</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Awasthi, Arun Kumar; Sylwester, Barbara; Sylwester, Janusz; Jain, Rajmal</p> <p>2016-06-01</p> <p>We investigate the evolution of the differential emission measure distribution (DEM[T]) in various phases of a B8.3 flare which occurred on 2009 July 04. We analyze the soft X-ray (SXR) emission in the 1.6-8.0 <span class="hlt">keV</span> range, recorded collectively by the Solar Photometer in X-rays (SphinX; Polish) and the Solar X-ray Spectrometer (Indian) instruments. We conduct a comparative investigation of the best-fit DEM[T] distributions derived by employing various inversion schemes, namely, single Gaussian, power-law functions and a Withbroe-Sylwester (W-S) maximum likelihood algorithm. In addition, the SXR spectrum in three different <span class="hlt">energy</span> <span class="hlt">bands</span>, that is, 1.6-5.0 <span class="hlt">keV</span> (low), 5.0-8.0 <span class="hlt">keV</span> (high), and 1.6-8.0 <span class="hlt">keV</span> (combined), is analyzed to determine the dependence of the best-fit DEM[T] distribution on the selection of the <span class="hlt">energy</span> interval. The evolution of the DEM[T] distribution, derived using a W-S algorithm, reveals multi-thermal plasma during the rise to the maximum phase of the flare, and isothermal plasma in the post-maximum phase of the flare. The thermal <span class="hlt">energy</span> content is estimated by considering the flare plasma to be (1) isothermal and (2) multi-thermal in nature. We find that the <span class="hlt">energy</span> content during the flare, estimated using the multi-thermal approach, is in good agreement with that derived using the isothermal assumption, except during the flare maximum. Furthermore, the (multi-) thermal <span class="hlt">energy</span> estimated while employing the low-<span class="hlt">energy</span> <span class="hlt">band</span> of the SXR spectrum results in higher values than that derived from the combined <span class="hlt">energy</span> <span class="hlt">band</span>. On the contrary, the analysis of the high-<span class="hlt">energy</span> <span class="hlt">band</span> of the SXR spectrum leads to lower thermal <span class="hlt">energy</span> than that estimated from the combined <span class="hlt">energy</span> <span class="hlt">band</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22666222-thermal-characteristics-differential-emission-measure-distribution-during-b8-flare-july','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22666222-thermal-characteristics-differential-emission-measure-distribution-during-b8-flare-july"><span>THERMAL CHARACTERISTICS AND THE DIFFERENTIAL EMISSION MEASURE DISTRIBUTION DURING A B8.3 FLARE ON 2009 JULY 4</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>Awasthi, Arun Kumar; Sylwester, Barbara; Sylwester, Janusz</p> <p></p> <p>We investigate the evolution of the differential emission measure distribution (DEM[ T ]) in various phases of a B8.3 flare which occurred on 2009 July 04. We analyze the soft X-ray (SXR) emission in the 1.6–8.0 <span class="hlt">keV</span> range, recorded collectively by the Solar Photometer in X-rays (SphinX; Polish) and the Solar X-ray Spectrometer (Indian) instruments. We conduct a comparative investigation of the best-fit DEM[ T ] distributions derived by employing various inversion schemes, namely, single Gaussian, power-law functions and a Withbroe–Sylwester (W–S) maximum likelihood algorithm. In addition, the SXR spectrum in three different <span class="hlt">energy</span> <span class="hlt">bands</span>, that is, 1.6–5.0 <span class="hlt">keV</span> (low),more » 5.0–8.0 <span class="hlt">keV</span> (high), and 1.6–8.0 <span class="hlt">keV</span> (combined), is analyzed to determine the dependence of the best-fit DEM[ T ] distribution on the selection of the <span class="hlt">energy</span> interval. The evolution of the DEM[ T ] distribution, derived using a W–S algorithm, reveals multi-thermal plasma during the rise to the maximum phase of the flare, and isothermal plasma in the post-maximum phase of the flare. The thermal <span class="hlt">energy</span> content is estimated by considering the flare plasma to be (1) isothermal and (2) multi-thermal in nature. We find that the <span class="hlt">energy</span> content during the flare, estimated using the multi-thermal approach, is in good agreement with that derived using the isothermal assumption, except during the flare maximum. Furthermore, the (multi-) thermal <span class="hlt">energy</span> estimated while employing the low-<span class="hlt">energy</span> <span class="hlt">band</span> of the SXR spectrum results in higher values than that derived from the combined <span class="hlt">energy</span> <span class="hlt">band</span>. On the contrary, the analysis of the high-<span class="hlt">energy</span> <span class="hlt">band</span> of the SXR spectrum leads to lower thermal <span class="hlt">energy</span> than that estimated from the combined <span class="hlt">energy</span> <span class="hlt">band</span>.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AGUFMSH21A2385W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFMSH21A2385W"><span>The <span class="hlt">Energy</span> Spectrum of Solar Energetic Electrons</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wang, L.; Yang, L.; Krucker, S.; Wimmer-Schweingruber, R. F.; Bale, S. D.</p> <p>2015-12-01</p> <p>Here we present a statistical survey of the <span class="hlt">energy</span> spectrum of solar energetic electron events (SEEs) observed by the WIND 3DP instrument from 1995 though 2014. For SEEs with the minimum <span class="hlt">energy</span> below 10 <span class="hlt">keV</span> and the maximum <span class="hlt">energy</span> above 100 <span class="hlt">keV</span>, ~85% (~2%) have a double-power-law <span class="hlt">energy</span> spectrum with a steepening (hardening) above the break <span class="hlt">energy</span>, while ~13% have a single-power-law <span class="hlt">energy</span> spectrum at all <span class="hlt">energies</span>. The average spectral index is ~2.4 below the <span class="hlt">energy</span> break and is ~4.0 above the <span class="hlt">energy</span> break. For SEEs detected only at <span class="hlt">energies</span> <10 <span class="hlt">keV</span> (>20 <span class="hlt">keV</span>), they generally show a single-power-law spectrum with the average index of ~3.0 (~3.3). The spectrum of SEEs detected only below 10 <span class="hlt">keV</span> appears to get harder with increasing solar activity, but the spectrum of SEEs with higher-<span class="hlt">energy</span> electrons shows no clear correlation with solar activity. We will also investigate whether the observed <span class="hlt">energy</span> spectrum of SEEs at 1 AU mainly reflects the electron acceleration at the Sun or the electron transport in the interplanetary medium.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JPhCS.912a2039M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JPhCS.912a2039M"><span>Electromagnetic Design of a Radiofrequency Cavity</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Montoya Soto, G. R.; Duarte Galvan, Carlos; Monzon, Ildefonso Leon; Podesta Lerma, Pedro Luis manuel; Valerio-Lizarraga, C. A.</p> <p>2017-10-01</p> <p>Electromagnetic and mechanical studies have been performed with the aim of build a RF cavity in the S-<span class="hlt">Band</span> (2998 MHz), the design takes into consideration the relativistic change in the electron velocity through the acceleration cavity. Four cavity cases were considered at different input <span class="hlt">energies</span>, 50 <span class="hlt">KeV</span>, 100 <span class="hlt">KeV</span>, 150 <span class="hlt">KeV</span>, with output <span class="hlt">energies</span> of 350 <span class="hlt">KeV</span>, the designs show good acceleration efficiency and beam coherence comparable to the one created in the cathode.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/4352889-isomer-rotational-bands-sup-os','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/4352889-isomer-rotational-bands-sup-os"><span>A 6.1 s isomer and rotational <span class="hlt">bands</span> in $sup 192$Os</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>Pakkanen, A.; Heikkinen, D.W.</p> <p></p> <p>A 6.1 plus or minus 0.2 s activity has been observed when natural Os targets were bombarded with 14.5 MeV neutrons. The activity is assigned to the decay of a high-spin isomer in /sup 192/Os at 2015.4 <span class="hlt">keV</span>, which is depopulated by M2 and E3 transitions. Singles and coincidence gamma -ray spectra have allowed the identification of seven new states in /sup 192/Os. Several of these levels have been placed in either the ground- state or gamma -vibrational <span class="hlt">bands</span>, which are strongly mixed. Excitation <span class="hlt">energies</span>, B(E2) ratios for these <span class="hlt">bands</span> are compared with different theoretical models. (auth) It is shownmore » that, if one uses single-particle <span class="hlt">energies</span> from experiment and a delta residual interaction, it is not possible to obtain the <span class="hlt">energy</span> of the giant dipole and spurious states of /sup 208/Pb, and at the same time obtain reasonable results for the low-lying two- particle spectra of /sup 210/Pb or /sup 210/Po. Related to the above problem, the isobaric analog state of /sup 208/Pb (in /sup 208/Bi) comes much too low in calculations using realistic interactions. It is noted that the above difficulties can be overcome, phenomenologically at least, by adding to the effective interaction some longrange repulsive components. The Bansal- French and the Schiffer interactions are examples of these; however, the dipole--dipole component of the Schiffer interaction gives much too large a splitting between the dipole state and spurious state. (auth)« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JQSRT.204..232B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JQSRT.204..232B"><span>M sub-shell X-ray fluorescence cross-section measurements for six elements in the range Z = 78-92 at tuned synchrotron photon <span class="hlt">energies</span> 5, 7 and 9 <span class="hlt">keV</span>.</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bansal, Himani; Tiwari, M. K.; Mittal, Raj</p> <p>2018-01-01</p> <p>M sub-shell X-ray fluorescence cross-sections of elements Pt, Au, Hg, Pb, Th and U have been measured with linearly polarized photon beams from Indus-II synchrotron source at Raja Ramanna Centre for Advanced Technology (RRCAT), India at tuned 5, 7 and 9 <span class="hlt">keV</span> <span class="hlt">energies</span> less than the L3 edge <span class="hlt">energy</span> of elements. Measurements at present <span class="hlt">energies</span> and elements are not available in literature. Therefore, measured cross-sections for Mξ, Mδ, Mα, Mβ, Mγ, Mm1 and Mm2 group of X-rays were compared with calculated theoretical values based upon Non Relativistic Hartree-Slater (NRHS) and relativistic Dirac-Fork (DF) and Dirac-Hartree-Slater (DHS) models. The measured cross-sections along with our earlier quoted measurements at 8 and 10 <span class="hlt">keV</span> by Kaur et al. [Nucl. Instrum. Meth. B, 2014; 320: 37] are found in good agreement with DF and DHS values around 20% deviations and are highly deviated from NRHS values. Most of the spots of observed high deviations in measured and theoretical cross-sections are found to coincide with the presence of crisscrosses/sharp variations in contributing physical parameters photo-ionization cross-sections σMi's and Coster-Kronig yields fij's with Zs.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22370584-high-energy-ray-imaging-pulsar-wind-nebula-msh-constraints-particle-acceleration-transport','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22370584-high-energy-ray-imaging-pulsar-wind-nebula-msh-constraints-particle-acceleration-transport"><span>High-<span class="hlt">energy</span> X-ray imaging of the pulsar wind nebula MSH 15–52: constraints on particle acceleration and transport</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>An, Hongjun; Kaspi, Victoria M.; Madsen, Kristin K.</p> <p>2014-10-01</p> <p>We present the first images of the pulsar wind nebula (PWN) MSH 15–52 in the hard X-ray <span class="hlt">band</span> (≳8 <span class="hlt">keV</span>), as measured with the Nuclear Spectroscopic Telescope Array (NuSTAR). Overall, the morphology of the PWN as measured by NuSTAR in the 3-7 <span class="hlt">keV</span> <span class="hlt">band</span> is similar to that seen in Chandra high-resolution imaging. However, the spatial extent decreases with <span class="hlt">energy</span>, which we attribute to synchrotron <span class="hlt">energy</span> losses as the particles move away from the shock. The hard-<span class="hlt">band</span> maps show a relative deficit of counts in the northern region toward the RCW 89 thermal remnant, with significant asymmetry. We find thatmore » the integrated PWN spectra measured with NuSTAR and Chandra suggest that there is a spectral break at 6 <span class="hlt">keV</span>, which may be explained by a break in the synchrotron-emitting electron distribution at ∼200 TeV and/or imperfect cross calibration. We also measure spatially resolved spectra, showing that the spectrum of the PWN softens away from the central pulsar B1509–58, and that there exists a roughly sinusoidal variation of spectral hardness in the azimuthal direction. We discuss the results using particle flow models. We find non-monotonic structure in the variation with distance of spectral hardness within 50'' of the pulsar moving in the jet direction, which may imply particle and magnetic-field compression by magnetic hoop stress as previously suggested for this source. We also present two-dimensional maps of spectral parameters and find an interesting shell-like structure in the N {sub H} map. We discuss possible origins of the shell-like structure and their implications.« 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/2018EL....12112001Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018EL....12112001Z"><span>Precise measurement of <span class="hlt">energy</span> of the first excited state of 115Sn (Eexc ≃ 497.3 <span class="hlt">keV</span>)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zheltonozhsky, V. A.; Savrasov, A. M.; Strilchuk, N. V.; Tretyak, V. I.</p> <p>2018-01-01</p> <p>Single beta decay of 115In to the first excited level of 115Sn (E\\text{exc}≃ 497.3 \\text{keV} ) is known as β-decay with the lowest Qβ value. To determine the Qβ precisely, one has to measure very accurately the E\\text{exc} value. A sample of tin enriched in 115Sn to 50.7% was irradiated by a proton beam at the U-120 accelerator of INR, Kyiv. The 115Sb radioactive isotope, created in the 115Sn(p,n)115Sb reaction, decays with T1/2 = 32 \\text{min} to 115Sn populating the 497 <span class="hlt">keV</span> level with ≃ 96{%} probability. The total statistics of ˜105 counts collected in the 497 <span class="hlt">keV</span> peak in a series of measurements, the exact description of the peak shape and the precisely known calibration points around the 497 <span class="hlt">keV</span> peak allowed to obtain the value E\\text{exc}= 497.342(3) \\text{keV} , which is the most precise to-date. This leads to the following Qβ\\ast value for the decay 115In → 115Sn*: Qβ\\ast= 147 +/- 10 \\text{eV} .</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017MNRAS.471..364B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017MNRAS.471..364B"><span>Inferring Compton-thick AGN candidates at z > 2 with Chandra using the >8 <span class="hlt">keV</span> rest-frame spectral curvature</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Baronchelli, L.; Koss, M.; Schawinski, K.; Cardamone, C.; Civano, F.; Comastri, A.; Elvis, M.; Lanzuisi, G.; Marchesi, S.; Ricci, C.; Salvato, M.; Trakhtenbrot, B.; Treister, E.</p> <p>2017-10-01</p> <p>To fully understand cosmic black hole growth, we need to constrain the population of heavily obscured active galactic nuclei (AGNs) at the peak of cosmic black hole growth (z ˜1-3). Sources with obscuring column densities higher than 1024 atoms cm-2, called Compton-thick (CT) AGNs, can be identified by excess X-ray emission at ˜20-30 <span class="hlt">keV</span>, called the 'Compton hump'. We apply the recently developed Spectral Curvature (SC) method to high-redshift AGNs (2 < z < 5) detected with Chandra. This method parametrizes the characteristic 'Compton hump' feature cosmologically redshifted into the X-ray <span class="hlt">band</span> at observed <span class="hlt">energies</span> <10 <span class="hlt">keV</span>. We find good agreement in CT AGNs found using the SC method, and bright sources fit using their full spectrum with X-ray spectroscopy. In the Chandra Deep Field-South, we measure a CT fraction of 17^{+19}_{-11} per cent (3/17) for sources with observed luminosity >5 × 1043erg s-1. In the Cosmological Evolution Survey (COSMOS), we find an observed CT fraction of 15^{+4}_{-3} per cent (40/272) or 32 ± 11 per cent when corrected for the survey sensitivity. When comparing to low redshift AGNs with similar X-ray luminosities, our results imply that the CT AGN fraction is consistent with having no redshift evolution. Finally, we provide SC equations that can be used to find high-redshift CT AGNs (z > 1) for current (XMM-Newton) and future (eROSITA and ATHENA) X-ray missions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26075571','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26075571"><span>Structure determination from XAFS using high-accuracy measurements of x-ray mass attenuation coefficients of silver, 11 <span class="hlt">keV</span>-28 <span class="hlt">keV</span>, and development of an all-<span class="hlt">energies</span> approach to local dynamical analysis of bond length, revealing variation of effective thermal contributions across the XAFS spectrum.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Tantau, L J; Chantler, C T; Bourke, J D; Islam, M T; Payne, A T; Rae, N A; Tran, C Q</p> <p>2015-07-08</p> <p>We use the x-ray extended range technique (XERT) to experimentally determine the mass attenuation coefficient of silver in the x-ray <span class="hlt">energy</span> range 11 <span class="hlt">kev</span>-28 <span class="hlt">kev</span> including the silver K absorption edge. The results are accurate to better than 0.1%, permitting critical tests of atomic and solid state theory. This is one of the most accurate demonstrations of cross-platform accuracy in synchrotron studies thus far. We derive the mass absorption coefficients and the imaginary component of the form factor over this range. We apply conventional XAFS analytic techniques, extended to include error propagation and uncertainty, yielding bond lengths accurate to approximately 0.24% and thermal Debye-Waller parameters accurate to 30%. We then introduce the FDMX technique for accurate analysis of such data across the full XAFS spectrum, built on full-potential theory, yielding a bond length accuracy of order 0.1% and the demonstration that a single Debye parameter is inadequate and inconsistent across the XAFS range. Two effective Debye-Waller parameters are determined: a high-<span class="hlt">energy</span> value based on the highly-correlated motion of bonded atoms (σ(DW) = 0.1413(21) Å), and an uncorrelated bulk value (σ(DW) = 0.1766(9) Å) in good agreement with that derived from (room-temperature) crystallography.</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/2015PASJ...67...41Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015PASJ...67...41Y"><span>Sub-MeV <span class="hlt">band</span> observation of a hard burst from AXP 1E 1547.0-5408 with the Suzaku Wide-<span class="hlt">band</span> All-sky Monitor</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yasuda, Tetsuya; Iwakiri, Wataru B.; Tashiro, Makoto S.; Terada, Yukikatsu; Kouzu, Tomomi; Enoto, Teruaki; Nakagawa, Yujin E.; Bamba, Aya; Urata, Yuji; Yamaoka, Kazutaka; Ohno, Masanori; Shibata, Shinpei; Makishima, Kazuo</p> <p>2015-06-01</p> <p>The 2.1-s anomalous X-ray pulsar 1E 1547.0-5408 exhibited an X-ray outburst on 2009 January 22, emitting a large number of short bursts. The wide-<span class="hlt">band</span> all-sky monitor (WAM) on-board Suzaku detected at least 254 bursts in the 160 <span class="hlt">keV</span>-6.2 MeV <span class="hlt">band</span> over the period of January 22 00:57-17:02 UT from the direction of 1E 1547.0-5408. One of these bursts, which occurred at 06:45:13, produced the brightest fluence in the 0.5-6.2 MeV range, with an averaged 0.16-6.2 MeV flux and extrapolated 25 <span class="hlt">keV</span>-2 MeV fluence of about 1 × 10-5 erg cm-2 s-1 and about 3 × 10-4 erg cm-2, respectively. After pile-up corrections, the time-resolved WAM spectra of this burst were well-fitted in the 0.16-6.2 MeV range by two-component models; specifically, a blackbody plus an optically thin thermal bremsstrahlung or a combination of a blackbody and a power-law component with an exponential cut-off. These results are compared with previous works reporting the persistent emission and weaker short bursts followed by the same outburst.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/5480774-preparation-relatively-clean-carbon-backings-used-charged-particle-induced-ray-studies-rays-below-kev','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/5480774-preparation-relatively-clean-carbon-backings-used-charged-particle-induced-ray-studies-rays-below-kev"><span>Preparation of relatively clean carbon backings used in charged particle induced x-ray studies for x-rays below 4 <span class="hlt">KeV</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>Kocur, P.; Duggan, J.L.; McDaniel, F.D.</p> <p>1983-04-01</p> <p>In a recent series of studies of M-shell ionization induced by protons, alpha particles, and fluorine ions, an unmanageable background of low <span class="hlt">energy</span> contaminant x rays was observed. These K-shell x rays were primarily from Ca, K, Cl, S, P, Si and Na. The <span class="hlt">energy</span> range of these contaminants is from 3.691 to 1.041 <span class="hlt">keV</span>. The M-shell x rays being studied were for various elements from U ( about 3.5 <span class="hlt">keV</span>) down to Eu (1.5 <span class="hlt">keV</span>). In order to evaluate and reduce the problem, the contaminants for carbon foils from a number of different manufacturers and a wide variety ofmore » foil float-off procedures have been studied. Carbon foils have been produced in our laboratory using carbon rods from several different manufacturers. In this paper, techniques will be described that are most appropriate to reduce the above contaminants to a reasonable level. These techniques should be useful in trace element analysis (PIXE) studies and fundamental ionization measurements for low x-ray <span class="hlt">energies</span>.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22409711-we-bre-impact-iudr-rat-glioma-cell-survival-kev-photo-activated-auger-electron-therapy','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22409711-we-bre-impact-iudr-rat-glioma-cell-survival-kev-photo-activated-auger-electron-therapy"><span>WE-E-BRE-08: Impact of IUdR in Rat 9L Glioma Cell Survival for 25–35 <span class="hlt">KeV</span> Photo-Activated Auger Electron Therapy</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>Alvarez, D; Hogstrom, K; Mary Bird Perkins Cancer Center, Baton Rouge, LA</p> <p>2014-06-15</p> <p>Purpose: To determine the biological effect from Auger electrons with 9% and 18% iododeoxyuridine (IUdR) incorporated into the DNA of rat 9L glioma cells at photon <span class="hlt">energies</span> above and below the K-edge of iodine (33.2 <span class="hlt">keV</span>). Methods: Rat 9L glioma cell survival versus dose curves with 0%, 9%, and 18% thymidine replacement with IUdR were measured using four irradiation <span class="hlt">energies</span> (4 MV x-rays; monochromatic 35, 30, and 25 <span class="hlt">keV</span> synchrotron photons). For each of 11 conditions (<span class="hlt">Energy</span>, %IUdR) survival curves were fit to the data (826 cell cultures) using the linear-quadratic model. The ratio of doses resulting in 10% survivalmore » gave sensitization enhancement ratios (SER10) from which contributions due to linear-<span class="hlt">energy</span> transfer (LET), radiosensitization (RS), and Auger effect (AE) were extracted. Results: At 35, 30, and 25 <span class="hlt">keV</span>, SER10,LET values were 1.08±0.03, 1.22±0.02, and 1.37±0.02, respectively. At 4 MV SER10,RS values for 9% and 18% IUdR were 1.28±0.02 and 1.40±0.02, respectively. Assuming LET effects are independent of %IUdR and radiosensitization effects are independent of <span class="hlt">energy</span>, SER10,AE values for 18% IUdR at 35, 30, and 25 <span class="hlt">keV</span> were 1.35±0.05, 1.06±0.03, and 0.98±0.03, respectively; values for 9% IUdR at 35 and 25 <span class="hlt">keV</span> were 1.01±0.04 and 0.82±0.02, respectively. Conclusion: For 18% IUdR the radiosensitization effect of 1.40 and the Auger effect of 1.35 at 35 <span class="hlt">keV</span> are equally important to the combined effect of 1.90. No measureable Auger effect was observed for <span class="hlt">energies</span> below the K-edge at 20 and 25 <span class="hlt">keV</span>, as expected. The insignificant Auger effect at 9% IUdR was not expected. Additional data (40–70 <span class="hlt">keV</span>) and radiobiological modeling are being acquired to better understand the <span class="hlt">energy</span> dependence of Auger electron therapy with IUdR. Funding support in part by the National Science Foundation Graduate Research Fellowship Program and in part by Contract No. W81XWH-10-1-0005 awarded by the U.S. Army Research Acquisition Activity. This paper does not</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('http://adsabs.harvard.edu/abs/2017EPJWC.14604042D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EPJWC.14604042D"><span>Fission-fragment total kinetic <span class="hlt">energy</span> and mass yields for neutron-induced fission of 235U and 238U with En =200 <span class="hlt">keV</span> - 30 MeV</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Duke, D. L.; Tovesson, F.; Brys, T.; Geppert-Kleinrath, V.; Hambsch, F.-J.; Laptev, A.; Meharchand, R.; Manning, B.; Mayorov, D.; Meierbachtol, K.; Mosby, S.; Perdue, B.; Richman, D.; Shields, D.; Vidali, M.</p> <p>2017-09-01</p> <p>The average Total Kinetic <span class="hlt">Energy</span> (TKE) release and fission-fragment yields in neutron-induced fission of 235U and 238U was measured using a Frisch-gridded ionization chamber. These observables are important nuclear data quantites that are relevant to applications and for informing the next generation of fission models. The measurements were performed a the Los Alamos Neutron Science Center and cover En = 200 <span class="hlt">keV</span> - 30 MeV. The double-<span class="hlt">energy</span> (2E) method was used to determine the fission-fragment yields and two methods of correcting for prompt-neutron emission were explored. The results of this study are correlated mass and TKE data.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010cosp...38.1594B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010cosp...38.1594B"><span>Energetic Neutral Atom Spectra in the 0.2-3.0 <span class="hlt">keV</span> from a Residual Source Across the Sky Obtained by the Neutral Particle Detector on board Venus Express</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Brandt, Pontus; Roelof, Edmond; Wurz, Peter; Decker, Robert; Barabash, Stas; Bazell, David; Sotirelis, Thomas</p> <p></p> <p>We have surveyed the sky for residual energetic neutral atom (ENA) signals in the <span class="hlt">energy</span> range of 0.2-3.0 <span class="hlt">keV</span> [Brandt et al., AIP Proceedings, 2009]. Approximately three years of data obtained by the Neutral Particle Detector (NPD) on board Venus Express (VEX) from May 2006 through August 2009 have been analyzed. After applying strict viewing criteria to minimize all known signals and subtracting the UV background from the Milky Way, we find a residual <span class="hlt">energy</span> spectral shape with a ledge/bump at around 0.5 <span class="hlt">keV</span> and a break in the spectral slope at about 1.0 <span class="hlt">keV</span>, reiminiscent of the spectral shape obtained in reverse shocks. The ledge/bump at about 0.5 <span class="hlt">keV</span> appears consistent with twice the plasma flow velocity obtained by the V1 measurements in the inner HS. When the ENA spectrum is divided by the <span class="hlt">energy</span> dependent charge exchange cross section its slope above 1 <span class="hlt">keV</span> has a spectral power-law index of 1.5, with some variations across the sky. In order to better understand the spectral shape over an extended <span class="hlt">energy</span> range we compare the spectra obtained by VEX/NPD with the ones reported by the Interstellar Boundary Explorer (IBEX) [Funsten et al., Science, 2009], by the Ion Neutral Camera (INCA) on board Cassini [Krimigis et al., Science, 2009]and with those measured in-situ in the inner heliosheath (HS) by the Low-<span class="hlt">Energy</span> Charged Particle (LECP) instrument (>40 <span class="hlt">keV</span>) on board Voyager-1 (V1).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015IAUGA..2254894A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015IAUGA..2254894A"><span>Thermal characteristics of multi-wavelength emission during a B8.3 flare occurred on July 04, 2009</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Awasthi, Arun Kumar; Sylwester, Barbara; Sylwester, Janusz; Jain, Rajmal</p> <p>2015-08-01</p> <p>We explore the temporal evolution of flare plasma parameters including temperature (T) - differential emission measure (DEM) relationship by analyzing high spectral and temporal cadence X-ray emission in 1.2-20 <span class="hlt">keV</span> <span class="hlt">energy</span> <span class="hlt">band</span>, recorded by SphinX (Polish) and Solar X-ray Spectrometer (SOXS; Indian) instruments, during a B8.3 flare which occurred on July 04, 2009. SphinX records X-ray emission in 1.2-15 <span class="hlt">keV</span> <span class="hlt">energy</span> <span class="hlt">band</span> with the temporal and spectral cadence as good as 6µs and 0.4 <span class="hlt">keV</span>, respectively. On the other hand, SOXS provides X-ray observations in 4-25 <span class="hlt">keV</span> <span class="hlt">energy</span> <span class="hlt">band</span> with the temporal and spectral resolution of 3s and 0.7 <span class="hlt">keV</span>, respectively. In addition, we integrate co-temporal EUV line emission in 171, 194 and 284 angstrom obtained from STEREO mission in order to explore low-temperature response to the flare emission. In order to fit observed evolution of multi-wavelength emission during the flare, we incorporate multi-Gaussian and well-established Withbroe - Sylwester maximum likelihood DEM inversion algorithms. Thermal energetics are also estimated using geometrically corrected flaring loop structure obtained through EUV images of the active region from STEREO twin satellites. In addition, we also study the trigger and <span class="hlt">energy</span> release scenario of this low-intensity class flare in terms of magnetic field as well as multi-wavelength emission.</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('http://adsabs.harvard.edu/abs/2009AIPC.1126..185B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009AIPC.1126..185B"><span>Black-hole Binaries: Life Begins at 40 <span class="hlt">keV</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Belloni, Tomaso M.; Motta, Sara</p> <p>2009-05-01</p> <p>In the study of black-hole transients, an important problem that still needs to be answered is how the high-<span class="hlt">energy</span> part of the spectrum evolves from the low-hard to the high-soft state, given that they have very different properties. Recent results obtained with RXTE and INTEGRAL have given inconsistent results. With RXTE, we have found that the high-<span class="hlt">energy</span> cutoff in GX 339-4 during the transition first decreases (during the low-hard state), then increases again across the Hard-Intermediate state, to become unmeasurable in the soft states (possibly because of statistical limitations). We show Simbol-X will be able to determine the spectral shape with superb accuracy. As the high-<span class="hlt">energy</span> part of the spectrum is relatively less known than the one below 20 <span class="hlt">keV</span>, Simbol-X will provide important results that will help out understanding of the extreme physical conditions in the vicinity of a stellar-mass black hole.</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/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 addition, 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/2009ApJ...696.1517K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009ApJ...696.1517K"><span>The Solar Wind Charge-Transfer X-Ray Emission in the 1/4 <span class="hlt">keV</span> <span class="hlt">Energy</span> Range: Inferences on Local Bubble Hot Gas at Low Z</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Koutroumpa, D.; Lallement, R.; Raymond, J. C.; Kharchenko, V.</p> <p>2009-05-01</p> <p>We present calculations of the heliospheric solar wind charge-exchange (SWCX) emission spectra and the resulting contributions of this diffuse background in the ROSAT 1/4 <span class="hlt">keV</span> <span class="hlt">bands</span>. We compare our results with the soft X-ray background (SXRB) emission detected in front of 378 identified shadowing regions during the ROSAT All-Sky Survey. This foreground component is principally attributed to the hot gas of the so-called Local Bubble (LB), an irregularly shaped cavity of ~50-150 pc around the Sun, which is supposed to contain ~106 K plasma. Our results suggest that the SWCX emission from the heliosphere is bright enough to account for most of the foreground emission toward the majority of low galactic latitude directions, where the LB is the least extended. On the other hand, in a large part of directions with galactic latitude above 30°, the heliospheric SWCX intensity is significantly smaller than the measured one. However, the SWCX R2/R1 <span class="hlt">band</span> ratio differs slightly from the data in the galactic center direction, and more significantly in the galactic anticentre direction where the observed ratio is the smallest. Assuming that both SWCX and hot gas emission are present and their relative contributions vary with direction, we tested a series of thermal plasma spectra for temperatures ranging from 10 5 to 10 6.5 K and searched for a combination of SWCX spectra and thermal emission matching the observed intensities and <span class="hlt">band</span> ratios, while simultaneously being compatible with O VI emission measurements. In the frame of collisional equilibrium models and for solar abundances, the range we derive for hot gas temperature and emission measure cannot reproduce the Wisconsin C/B <span class="hlt">band</span> ratio. This implies that accounting for SWCX contamination does not remove these known disagreements between data and classical hot gas models. We emphasize the need for additional atomic data, describing consistently EUV and X-ray photon spectra of the charge-exchange emission of heavier solar</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015yCat..35760126A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015yCat..35760126A"><span>VizieR Online Data Catalog: The 2009 multiwavelength campaign on Mrk421 (Aleksic+, 2015)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Aleksic, J.; Ansoldi, S.; Antonelli, L. A.; Antoranz, P.; Babic, A.; Bangale, P.; Barres de Almeida, U.; Barrio, J. A.; Becerra Gonzalez, J.; Bednarek, W.; Berger, K.; Bernardini, E.; Biland, A.; Blanch, O.; Bock, R. K.; Bonnefoy, S.; Bonnoli, G.; Borracci, F.; Bretz, T.; Carmona, E.; Carosi, A.; Carreto Fidalgo, D.; Colin, P.; Colombo, E.; Contreras, J. L.; Cortina, J.; Covino, S.; da Vela, P.; Dazzi, F.; de Angelis, A.; de Caneva, G.; de Lotto, B.; Delgado Mendez, C.; Doert, M.; Dominguez, A.; Dominis Prester, D.; Dorner, D.; Doro, M.; Einecke, S.; Eisenacher, D.; Elsaesser, D.; Farina, E.; Ferenc, D.; Fonseca, M. V.; Font, L.; Frantzen, K.; Fruck, C.; Garcia Lopez, R. J.; Garczarczyk, M.; Garrido Terrats, D.; Gaug, M.; Giavitto, G.; Godinovi, N.; Gonzalez Munoz, A.; Gozzini, S. R.; Hadamek, A.; Hadasch, D.; Herrero, A.; Hildebrand, D.; Hose, J.; Hrupec, D.; Idec, W.; Kadenius, V.; Kellermann, H.; Knoetig, M. L.; Krause, J.; Kushida, J.; La Barbera, A.; Lelas, D.; Lewandowska, N.; Lindfors, E.; Longo, F.; Lombardi, S.; Lopez, M.; Lopez-Coto, R.; Lopez-Oramas, A.; Lorenz, E.; Lozano, I.; Makariev, M.; Mallot, K.; Maneva, G.; Mankuzhiyil, N.; Mannheim, K.; Maraschi, L.; Marcote, B.; Mariotti, M.; Martinez, M.; Mazin, D.; Menzel, U.; Meucci, M.; Miranda, J. M.; Mirzoyan, R.; Moralejo, A.; Munar-Adrover, P.; Nakajima, D.; Niedzwiecki, A.; Nilsson, K.; Nowak, N.; Orito, R.; Overkemping, A.; Paiano, S.; Palatiello, M.; Paneque, D.; Paoletti, R.; Paredes, J. M.; Paredes-Fortuny, X.; Partini, S.; Persic, M.; Prada, F.; Prada Moroni, P. G.; Prandini, E.; Preziuso, S.; Puljak, I.; Reinthal, R.; Rhode, W.; Ribo, M.; Rico, J.; Rodriguez Garcia, J.; Rugamer, S.; Saggion, A.; Saito, K.; Salvati, M.; Satalecka, K.; Scalzotto, V.; Scapin, V.; Schultz, C.; Schweizer, T.; Shore, S. N.; Sillanpaa, A.; Sitarek, J.; Snidaric, I.; Sobczynska, D.; Spanier, F.; Stamatescu, V.; Stamerra, A.; Steinbring, T.; Storz, J.; Sun, S.; Suri, T.; Takalo, L.; Tavecchio, F.; Temnikov, P.; Terzi, T.; Tescaro, D.; Teshima, M.; Thaele, J.; Tibolla, O.; Torres, D. F.; Toyama, T.; Treves, A.; Uellenbeck, M.; Vogler, P.; Wagner, R. M.; Zandanel, F.; Zanin, R.</p> <p>2015-02-01</p> <p>Light Curves of Mrk 421 as shown in Figure 1 of the paper are presented for following instruments/observatories and <span class="hlt">bands</span> (radio to very high <span class="hlt">energy</span> gamma rays): Effelsberg (2.6, 4.6, 7.9, 10.3, 13.6, 21.7, 31GHz), Medicina (8.4GHz), Metsahovi (37GHz), OVRO (15GHz), Noto (8, 22GHz), OAGH (J, H, K <span class="hlt">bands</span>), WIRO (J, K <span class="hlt">bands</span>), MITSuME (g, Rc, Ic <span class="hlt">bands</span>), ROVOR (B <span class="hlt">band</span>), GRT (V, R, B, I <span class="hlt">bands</span>), GASP (R <span class="hlt">band</span>), Steward (V <span class="hlt">band</span>), Swift/UVOT (UVW1, UVM2, UVW2), Swift/XRT (0.3-2 and 2-10<span class="hlt">keV</span>), RXTE/PCA (2-10<span class="hlt">keV</span>). RXTE/ASM (2-10<span class="hlt">keV</span>), Swift/BAT (15-50<span class="hlt">keV</span>), Fermi-LAT (>0.3<span class="hlt">keV</span>), Whipple (>300GeV), MAGIC (>300GeV). The observation period is from 2009 January 19 (MJD 54850) to 2009 June 1st (MJD 54983), where Mrk 421 was observed approximately once every two days. The Fermi-LAT photon fluxes are integrated over a three-day-long time interval, the RXTE/ASM and Swift/BAT photon fluxes over a seven-day long time interval. The Whipple 10-meter data (with an <span class="hlt">energy</span> threshold of 400GeV) were converted into fluxes above 300GeV using a power-law spectrum with index of 2.5. Host galaxy fluxes are given where a good estimate is available, which is the case for some optical <span class="hlt">bands</span> only. In the infrared, e.g., the host galaxy flux is larger than in the R <span class="hlt">band</span>, however, we do not have a good estimate of the galaxy flux and therefore it is not given in the table. (1 data file).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19950040034&hterms=SMM&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3DSMM','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19950040034&hterms=SMM&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3DSMM"><span>SMM observations of gamma-ray transients. 2: A search for gamma-ray lines between 400 and 600 <span class="hlt">keV</span> from the Crab Nebula</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Harris, Michael J.; Share, Gerald H.; Leising, Mark D.</p> <p>1994-01-01</p> <p>We have search spectra obtained by the Solar Maximum Mission Gamma-Ray Spectrometer during 1981-1988 for evidence of transient gamma-ray lines from the Crab Nebula which have been reported by previous experiments at <span class="hlt">energies</span> 400-460 <span class="hlt">keV</span> and 539 <span class="hlt">keV</span>. We find no evidence for significant emission in any of these lines on time scales between aproximately 1 day and approximately 1 yr. Our 3 sigma upper limits on the transient flux during 1 d intervals are approximately equal to 2.2 x 10(exp -3) photons/sq cm/s for narrow lines at any <span class="hlt">energy</span>, and approximately equal to 2.9 x 10(exp -3) photons/sq cm/s for the 539 <span class="hlt">keV</span> line if it is as broad as 42 <span class="hlt">keV</span> Full Width at Half Maximum (FWHM). We also searched our data during the approximately 5 hr period on 1981 June 6 during which Owens, Myers, & Thompson (1985) reported a strong line at 405 <span class="hlt">keV</span>. We detected no line down to a 3 upper sigma limit of 3.3 x 10(exp -3) photons/sq cm/s in disagreement with the flux 7.2 +/- 2.1 x 10(exp -3) photos/sq cm/s measured by Owens et al.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016HEAD...1511601O','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016HEAD...1511601O"><span>The Transient High <span class="hlt">Energy</span> Sky and Early Universe Surveyor</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>O'Brien, P. T.</p> <p>2016-04-01</p> <p>The Transient High <span class="hlt">Energy</span> Sky and Early Universe Surveyor is a mission which will be proposed for the ESA M5 call. THESEUS will address multiple components in the Early Universe ESA Cosmic Vision theme:4.1 Early Universe,4.2 The Universe taking shape, and4.3 The evolving violent Universe.THESEUS aims at vastly increasing the discovery space of the high <span class="hlt">energy</span> transient phenomena over the entire cosmic history. This is achieved via a unique payload providing an unprecedented combination of: (i) wide and deep sky monitoring in a broad <span class="hlt">energy</span> <span class="hlt">band</span>(0.3 <span class="hlt">keV</span>-20 MeV; (ii) focusing capabilities in the soft X-ray <span class="hlt">band</span> granting large grasp and high angular resolution; and (iii) on board near-IR capabilities for immediate transient identification and first redshift estimate.The THESEUS payload consists of: (i) the Soft X--ray Imager (SXI), a set of Lobster Eye (0.3--6 <span class="hlt">keV</span>) telescopes with CCD detectors covering a total FOV of 1 sr; (ii) the X--Gamma-rays spectrometer (XGS), a non-imaging spectrometer (XGS) based on SDD+CsI, covering the same FOV than the Lobster telescope extending the THESEUS <span class="hlt">energy</span> <span class="hlt">band</span> up to 20 MeV; and (iii) a 70cm class InfraRed Telescope (IRT) observing up to 2 microns with imaging and moderate spectral capabilities.The main scientific goals of THESEUS are to:(a) Explore the Early Universe (cosmic dawn and reionization era) by unveiling the Gamma--Ray Burst (GRBs) population in the first billion years}, determining when did the first stars form, and investigating the re-ionization epoch, the interstellar medium (ISM) and the intergalactic medium (IGM) at high redshifts.(b) Perform an unprecedented deep survey of the soft X-ray transient Universe in order to fill the present gap in the discovery space of new classes of transient; provide a fundamental step forward in the comprehension of the physics of various classes of Galactic and extra--Galactic transients, and provide real time trigger and accurate locations of transients for follow-up with next</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/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/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('https://www.ncbi.nlm.nih.gov/pubmed/22889459','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22889459"><span>Single-source dual-<span class="hlt">energy</span> spectral multidetector CT of pancreatic adenocarcinoma: optimization of <span class="hlt">energy</span> level viewing significantly increases lesion contrast.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Patel, B N; Thomas, J V; Lockhart, M E; Berland, L L; Morgan, D E</p> <p>2013-02-01</p> <p>To evaluate lesion contrast in pancreatic adenocarcinoma patients using spectral multidetector computed tomography (MDCT) analysis. The present institutional review board-approved, Health Insurance Portability and Accountability Act of 1996 (HIPAA)-compliant retrospective study evaluated 64 consecutive adults with pancreatic adenocarcinoma examined using a standardized, multiphasic protocol on a single-source, dual-<span class="hlt">energy</span> MDCT system. Pancreatic phase images (35 s) were acquired in dual-<span class="hlt">energy</span> mode; unenhanced and portal venous phases used standard MDCT. Lesion contrast was evaluated on an independent workstation using dual-<span class="hlt">energy</span> analysis software, comparing tumour to non-tumoural pancreas attenuation (HU) differences and tumour diameter at three <span class="hlt">energy</span> levels: 70 <span class="hlt">keV</span>; individual subject-optimized viewing <span class="hlt">energy</span> level (based on the maximum contrast-to-noise ratio, CNR); and 45 <span class="hlt">keV</span>. The image noise was measured for the same three <span class="hlt">energies</span>. Differences in lesion contrast, diameter, and noise between the different <span class="hlt">energy</span> levels were analysed using analysis of variance (ANOVA). Quantitative differences in contrast gain between 70 <span class="hlt">keV</span> and CNR-optimized viewing <span class="hlt">energies</span>, and between CNR-optimized and 45 <span class="hlt">keV</span> were compared using the paired t-test. Thirty-four women and 30 men (mean age 68 years) had a mean tumour diameter of 3.6 cm. The median optimized <span class="hlt">energy</span> level was 50 <span class="hlt">keV</span> (range 40-77). The mean ± SD lesion contrast values (non-tumoural pancreas - tumour attenuation) were: 57 ± 29, 115 ± 70, and 146 ± 74 HU (p = 0.0005); the lengths of the tumours were: 3.6, 3.3, and 3.1 cm, respectively (p = 0.026); and the contrast to noise ratios were: 24 ± 7, 39 ± 12, and 59 ± 17 (p = 0.0005) for 70 <span class="hlt">keV</span>, the optimized <span class="hlt">energy</span> level, and 45 <span class="hlt">keV</span>, respectively. For individuals, the mean ± SD contrast gain from 70 <span class="hlt">keV</span> to the optimized <span class="hlt">energy</span> level was 59 ± 45 HU; and the mean ± SD contrast gain from the optimized <span class="hlt">energy</span> level to 45 <span class="hlt">keV</span> was 31 ± 25 HU (p = 0</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 addition 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> <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://hdl.handle.net/2060/19990105712','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19990105712"><span>XTE Proposal #20102--"SS 433's High <span class="hlt">Energy</span> Spectrum"</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Band, David L.; Blanco, P.; Rothschild, R.; Kawai, N.; Kotani, T.; Oka, T.; Wagner, R. M.; Hjellming, R.; Rupen, M.; Brinkmann, W.</p> <p>1999-01-01</p> <p>We observed the jet-producing compact binary system SS 433 with RXTE during three multiwavelength campaigns, the first in conjunction with ASCA observations, the second simultaneous with a VLA-VLBA-MERLIN campaign, and the third associated with a Nobeyama millimeter-<span class="hlt">band</span> campaign. All these campaigns included optical observations. Occurring at different jet precession and binary phases, the observations also monitored the system during a radio flare. The data provide SS 433's X-ray spectrum over more than an <span class="hlt">energy</span> decade, and track the spectral variations as the X-ray source was partially eclipsed. The continuum can be modeled as a power law with an exponential cutoff, which can be detected to approximately 50 <span class="hlt">keV</span>. Strong line emission is evident in the 5-10 <span class="hlt">keV</span> range which can be modeled as a broad line whose <span class="hlt">energy</span> is precession independent and a narrow line whose <span class="hlt">energy</span> does vary with jet precession phase; this line model is clearly an over simplification since the PCA does not have sufficient <span class="hlt">energy</span> resolution to detect the lines ASCA observed. The eclipses are deeper at high <span class="hlt">energy</span> and at jet precession phases when the jets are more inclined towards and away from us. A large radio flare occurred between two sets of X-ray monitoring observations; an X-ray observation at the peak of the flare found a softer spectrum with a flux approximately 1/3 that of the quiescent level.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20170003474&hterms=Symbiotic&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D90%26Ntt%3DSymbiotic','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20170003474&hterms=Symbiotic&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D90%26Ntt%3DSymbiotic"><span>Identification of the Hard X-Ray Source Dominating the E > 25 <span class="hlt">keV</span> Emission of the Nearby Galaxy M31</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Yukita, M.; Ptak, A.; Hornschemeier, A. E.; Wik, D.; Maccarone, T. J.; Pottschmidt, K.; Zezas, A.; Antoniou, V.; Ballhausen, R.; Lehmer, B. D.; <a style="text-decoration: none; " href="javascript:void(0); " onClick="displayelement('author_20170003474'); toggleEditAbsImage('author_20170003474_show'); toggleEditAbsImage('author_20170003474_hide'); "> <img style="display:inline; width:12px; height:12px; " src="images/arrow-up.gif" width="12" height="12" border="0" alt="hide" id="author_20170003474_show"> <img style="width:12px; height:12px; display:none; " src="images/arrow-down.gif" width="12" height="12" border="0" alt="hide" id="author_20170003474_hide"></p> <p>2017-01-01</p> <p>We report the identification of a bright hard X-ray source dominating the M31 bulge above 25 <span class="hlt">keV</span> from a simultaneous NuSTAR-Swift observation. We find that this source is the counterpart to Swift J0042.6+4112, which was previously detected in the Swift BAT All-Sky Hard X-Ray Survey. This Swift BAT source had been suggested to be the combined emission from a number of point sources; our new observations have identified a single X-ray source from 0.5 to 50 <span class="hlt">keV</span> as the counterpart for the first time. In the 0.5-10 <span class="hlt">keV</span> <span class="hlt">band</span>, the source had been classified as an X-ray Binary candidate in various Chandra and XMM-Newton studies; however, since it was not clearly associated with Swift J0042.6+4112, the previous E is less than 10keVobservations did not generate much attention. This source has a spectrum with a soft X-ray excess (kT approximately equal to 0.2 <span class="hlt">keV</span>) plus a hard spectrum with a power law of gamma approximately equal to 1 and a cutoff around 15-20 <span class="hlt">keV</span>, typical of the spectral characteristics of accreting pulsars. Unfortunately, any potential pulsation was undetected in the NuSTAR data, possibly due to insufficient photon statistics. The existing deep HST (Hubble Space Telescope) images exclude high-mass (greater than 3 times the radius of the moon) donors at the location of this source. The best interpretation for the nature of this source is an X-ray pulsar with an intermediate-mass (less than 3 times the radius of the moon M) companion or a symbiotic X-ray binary. We discuss other possibilities in more detail.</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('https://www.osti.gov/servlets/purl/1402659','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1402659"><span>Effects of ULF waves on local and global energetic particles: Particle <span class="hlt">energy</span> and species dependences</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>Li, L. Y.; Yu, J.; Cao, J. B.</p> <p></p> <p>After 06:13 UT on 24 August 2005, an interplanetary shock triggers large-amplitude ultralow-frequency (ULF) waves (|δB| ≥ 15 nT) in the Pc4–Pc5 wave <span class="hlt">band</span> (1.6–9 mHz) near the noon geosynchronous orbit (6.6 RE). The local and global effects of ULF waves on energetic particles are observed by five Los Alamos National Laboratory satellites at different magnetic local times. The large-amplitude ULF waves cause the synchronous oscillations of energetic electrons and protons (≥75 <span class="hlt">keV</span>) at the noon geosynchronous orbit. When the energetic particles have a negative phase space density radial gradient, they undergo rapid outward radial diffusion and loss in themore » wave activity region. In the particle drift paths without strong ULF waves, only the rapidly drifting energetic electrons (≥225 <span class="hlt">keV</span>) display <span class="hlt">energy</span>-dispersive oscillations and flux decays, whereas the slowly drifting electrons (<225 <span class="hlt">keV</span>) and protons (75–400 <span class="hlt">keV</span>) have no ULF oscillation and loss feature. When the dayside magnetopause is compressed to the geosynchronous orbit, most of energetic electrons and protons are rapidly lost because of open drift trajectories. Furthermore, the global and multicomposition particle measurements demonstrate that the effect of ULF waves on nonlocal particle flux depends on the particle <span class="hlt">energy</span> and species, whereas magnetopause shadowing effect is independent of the energetic particle species. For the rapidly drifting outer radiation belt particles (≥225 <span class="hlt">keV</span>), nonlocal particle loss/acceleration processes could also change their fluxes in the entire drift trajectory in the absence of “ Dst effect” and substorm injection.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1402659-effects-ulf-waves-local-global-energetic-particles-particle-energy-species-dependences','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1402659-effects-ulf-waves-local-global-energetic-particles-particle-energy-species-dependences"><span>Effects of ULF waves on local and global energetic particles: Particle <span class="hlt">energy</span> and species dependences</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Li, L. Y.; Yu, J.; Cao, J. B.; ...</p> <p>2016-11-05</p> <p>After 06:13 UT on 24 August 2005, an interplanetary shock triggers large-amplitude ultralow-frequency (ULF) waves (|δB| ≥ 15 nT) in the Pc4–Pc5 wave <span class="hlt">band</span> (1.6–9 mHz) near the noon geosynchronous orbit (6.6 RE). The local and global effects of ULF waves on energetic particles are observed by five Los Alamos National Laboratory satellites at different magnetic local times. The large-amplitude ULF waves cause the synchronous oscillations of energetic electrons and protons (≥75 <span class="hlt">keV</span>) at the noon geosynchronous orbit. When the energetic particles have a negative phase space density radial gradient, they undergo rapid outward radial diffusion and loss in themore » wave activity region. In the particle drift paths without strong ULF waves, only the rapidly drifting energetic electrons (≥225 <span class="hlt">keV</span>) display <span class="hlt">energy</span>-dispersive oscillations and flux decays, whereas the slowly drifting electrons (<225 <span class="hlt">keV</span>) and protons (75–400 <span class="hlt">keV</span>) have no ULF oscillation and loss feature. When the dayside magnetopause is compressed to the geosynchronous orbit, most of energetic electrons and protons are rapidly lost because of open drift trajectories. Furthermore, the global and multicomposition particle measurements demonstrate that the effect of ULF waves on nonlocal particle flux depends on the particle <span class="hlt">energy</span> and species, whereas magnetopause shadowing effect is independent of the energetic particle species. For the rapidly drifting outer radiation belt particles (≥225 <span class="hlt">keV</span>), nonlocal particle loss/acceleration processes could also change their fluxes in the entire drift trajectory in the absence of “ Dst effect” and substorm injection.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017MS%26E..168a2009I','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017MS%26E..168a2009I"><span>The formation of nanopores in metal materials after irradiation by beams of Ar+ with <span class="hlt">energy</span> of 30 <span class="hlt">keV</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ivchenko, V. A.</p> <p>2017-01-01</p> <p>In this paper are the results of direction observations of nanopores in the subsurface volume of metals materials Pt and Pd(CuAg) using field-ion microscopy (FIM). Radiation of tip specimens was carried out with ions having an <span class="hlt">energy</span> ˜ 25-30 <span class="hlt">keV</span> in the fluency range of 1016 - 1018 ions/cm2, the current density lying within 150- 340 µA/cm2. Nanopores have been observed immediately after removal of the first atomic layers from the irradiated surface. It was established that, the threshold for ion-implanted platinum corresponds to fluence F = 1017 ions/cm2. For Pd(CuAg) it was revealed that nanopores have been down to 80 nm deep with current density 340 µA/cm2. Their dimensions and volume fractions were determined. The obtained results can be used for prediction of radiation stability of materials based on fcc metals.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016Metro..53....7M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016Metro..53....7M"><span>High accuracy experimental determination of copper and zinc mass attenuation coefficients in the 100 eV to 30 <span class="hlt">keV</span> photon <span class="hlt">energy</span> range</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ménesguen, Y.; Gerlach, M.; Pollakowski, B.; Unterumsberger, R.; Haschke, M.; Beckhoff, B.; Lépy, M.-C.</p> <p>2016-02-01</p> <p>The knowledge of atomic fundamental parameters such as mass attenuation coefficients with low uncertainties, is of decisive importance in elemental quantification using x-ray fluorescence analysis techniques. Several databases are accessible and frequently used within a large community of users. These compilations are most often in good agreement for photon <span class="hlt">energies</span> in the hard x-ray ranges. However, they significantly differ for low photon <span class="hlt">energies</span> and around the absorption edges of any element. In a joint cooperation of the metrology institutes of France and Germany, mass attenuation coefficients of copper and zinc were determined experimentally in the photon <span class="hlt">energy</span> range from 100 eV to 30 <span class="hlt">keV</span> by independent approaches using monochromatized synchrotron radiation at SOLEIL (France) and BESSY II (Germany), respectively. The application of high-accuracy experimental techniques resulted in mass attenuation coefficient datasets determined with low uncertainties that are directly compared to existing databases. The novel datasets are expected to enhance the reliability of mass attenuation coefficients.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JInst..12C1015M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JInst..12C1015M"><span>Using the Medipix3 detector for direct electron imaging in the range 60 <span class="hlt">keV</span> to 200 <span class="hlt">keV</span> in electron microscopy</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mir, J. A.; Plackett, R.; Shipsey, I.; dos Santos, J. M. F.</p> <p>2017-11-01</p> <p>Hybrid pixel sensor technology such as the Medipix3 represents a unique tool for electron imaging. We have investigated its performance as a direct imaging detector using a Transmission Electron Microscope (TEM) which incorporated a Medipix3 detector with a 300 μm thick silicon layer compromising of 256×256 pixels at 55 μm pixel pitch. We present results taken with the Medipix3 in Single Pixel Mode (SPM) with electron beam <span class="hlt">energies</span> in the range, 60-200 <span class="hlt">keV</span> . Measurements of the Modulation Transfer Function (MTF) and the Detective Quantum Efficiency (DQE) were investigated. At a given beam <span class="hlt">energy</span>, the MTF data was acquired by deploying the established knife edge technique. Similarly, the experimental data required to determine DQE was obtained by acquiring a stack of images of a focused beam and of free space (flatfield) to determine the Noise Power Spectrum (NPS).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JInst..13R1001M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JInst..13R1001M"><span>Retraction: Using the Medipix3 detector for direct electron imaging in the range 60 <span class="hlt">keV</span> to 200 <span class="hlt">keV</span> in electron microscopy Retraction: Using the Medipix3 detector for direct electron imaging in the range 60 <span class="hlt">keV</span> to 200 <span class="hlt">keV</span> in electron microscopy</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mir, J. A.; Plackett, R.; Shipsey, I.; dos Santos, J. M. F.</p> <p>2018-01-01</p> <p>The paper "Using the Medipix3 detector for direct electron imaging in the range 60<span class="hlt">keV</span> to 200<span class="hlt">keV</span> in electron microscopy" by J.A. Mir, R. Plackett, I. Shipsey and J.M.F. dos Santos has been retracted following the authors' request on the basis of the existence of a disagreement about the ownership of the data, to prevent conflict between collaborators.</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.osti.gov/servlets/purl/863711','DOE-PATENT-XML'); return false;" href="https://www.osti.gov/servlets/purl/863711"><span>Compact, maintainable 80-<span class="hlt">KeV</span> neutral beam module</span></a></p> <p><a target="_blank" href="http://www.osti.gov/doepatents">DOEpatents</a></p> <p>Fink, Joel H.; Molvik, Arthur W.</p> <p>1980-01-01</p> <p>A compact, maintainable 80-<span class="hlt">keV</span> arc chamber, extractor module for a neutral beam system immersed in a vacuum of <10.sup.-2 Torr, incorporating a nested 60-<span class="hlt">keV</span> gradient shield located midway between the high voltage ion source and surrounding grounded frame. The shield reduces breakdown or arcing path length without increasing the voltage gradient, tends to keep electric fields normal to conducting surfaces rather than skewed and reduces the peak electric field around irregularities on the 80-<span class="hlt">keV</span> electrodes. The arc chamber or ion source is mounted separately from the extractor or ion accelerator to reduce misalignment of the accelerator and to permit separate maintenance to be performed on these systems. The separate mounting of the ion source provides for maintaining same without removing the ion accelerator.</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('http://adsabs.harvard.edu/abs/2004RScI...75.2056M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2004RScI...75.2056M"><span>Liquid helium cryostat with internal fluorescence detection for x-ray absorption studies in the 2-6 <span class="hlt">keV</span> <span class="hlt">energy</span> region</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>McFarlane Holman, Karen L.; Latimer, Matthew J.; Yachandra, Vittal K.</p> <p>2004-06-01</p> <p>X-ray absorption spectroscopy (XAS) in the intermediate x-ray region (2-6 <span class="hlt">keV</span>) for dilute biological samples has been limited because of detector/flux limitations and inadequate cryogenic instrumentation. We have designed and constructed a new tailpiece/sample chamber for a commercially available liquid helium cooled cryostat which overcomes difficulties related to low fluorescence signals by using thin window materials and incorporating an internal photodiode detector. With the apparatus, XAS data at the Cl, S, and Ca K edges have been collected on frozen solutions and biological samples at temperatures down to 60 K. A separate chamber has been incorporated for collecting room-temperature spectra of standard compounds (for <span class="hlt">energy</span> calibration purposes) which prevents contamination of the cryostat chamber and allows the sample to remain undisturbed, both important concerns for studying dilute and radiation-sensitive samples.</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/2017PhDT.......100T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017PhDT.......100T"><span>Synthesis of Germanium-Tin Alloys by Ion Implantation and Pulsed Laser Melting: Towards a Group IV 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>Tran, Tuan Thien</p> <p></p> <p> increase the retained Sn concentration up to 15 at.% after implantation, as well as significantly improving the crystal quality of the Ge-Sn layer after PLM. With the use of the capping layer, a good quality Ge-Sn layer with 9 at.% Sn has been achieved using Sn implants at an <span class="hlt">energy</span> of 120 <span class="hlt">keV</span>. However, the thin film alloys produced by 100 <span class="hlt">keV</span> or 120 <span class="hlt">keV</span> Sn implantation and PLM are shown to contain compressive strain as a result of the large lattice mismatch between Ge and high Sn content alloys. Such strain compromises the tendency towards a direct bandgap material and hence strain relaxation is highly desirable. A thermal stability study showed that the thin film strained material is metastable up to 400°C, but thereafter Sn comes out of solution and diffuses to the material surface. To investigate a possible pathway to the synthesis of strain-relaxed material, a higher Sn implant <span class="hlt">energy</span> of 350 <span class="hlt">keV</span> was used to produce thicker alloy layers. XRD/reciprocal space mapping showed that this thicker alloy material is largely relaxed after PLM, which is beneficial for the direct <span class="hlt">band</span> gap transition and solves the trade-off between higher Sn concentration and compressive strain. However, RBS indicates a sub-surface <span class="hlt">band</span> of disorder which suggested a possible mechanism for the strain relaxation. Indeed, TEM examination of such material showed the material relaxed via the generation of non-equilibrium threading defects. Despite such defects, a PL study of this relaxed material found photon emission at a wavelength of 2150 nm for 6-9 at.% Sn alloys. However, the intensity of the emission was variable across different Sn content alloys, presumably as a result of the threading defects. A possible pathway to removing such defects is given that may enable both photodetectors and lasers to be fabricated at wavelengths above 2mum.</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> <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 additional 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://ntrs.nasa.gov/search.jsp?R=19860042897&hterms=compound+microscope&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dcompound%2Bmicroscope','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19860042897&hterms=compound+microscope&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dcompound%2Bmicroscope"><span>Low <span class="hlt">energy</span> X-ray spectra measured with a mercuric iodide <span class="hlt">energy</span> dispersive spectrometer in a scanning electron microscope</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Iwanczyk, J. S.; Dabrowski, A. J.; Huth, G. C.; Bradley, J. G.; Conley, J. M.</p> <p>1986-01-01</p> <p>A mercuric iodide <span class="hlt">energy</span> dispersive X-ray spectrometer, with Peltier cooling provided for the detector and input field effect transistor, has been developed and tested in a scanning electron microscope. X-ray spectra were obtained with the 15 <span class="hlt">keV</span> electron beam. An <span class="hlt">energy</span> resolution of 225 eV (FWHM) for Mn-K(alpha) at 5.9 <span class="hlt">keV</span> and 195 eV (FWHM) for the Mg-K line at 1.25 <span class="hlt">keV</span> has been measured. Overall system noise level was 175 eV (FWHM). The detector system characterization with a carbon target demonstrated good <span class="hlt">energy</span> sensitivity at low <span class="hlt">energies</span> and lack of significant spectral artifacts at higher <span class="hlt">energies</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018MNRAS.475.1238K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018MNRAS.475.1238K"><span>The Fermi-LAT detection of magnetar-like pulsar PSR J1846-0258 at high-<span class="hlt">energy</span> gamma-rays</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kuiper, L.; Hermsen, W.; Dekker, A.</p> <p>2018-03-01</p> <p>We report the detection of the pulsed signal of the radio-quiet magnetar-like pulsar PSR J1846-0258 in the high-<span class="hlt">energy</span> γ-ray data of the Fermi Large Area Telescope (Fermi LAT). We produced phase-coherent timing models exploiting RXTE PCA and Swift XRT monitoring data for the post- (magnetar-like) outburst period from 2007 August 28 to 2016 September 4, with independent verification using INTEGRAL ISGRI and Fermi GBM data. Phase-folding barycentric arrival times of selected Fermi LAT events from PSR J1846-0258 resulted in a 4.2σ detection (30-100 MeV) of a broad pulse consistent in shape and aligned in phase with the profiles that we measured with Swift XRT (2.5-10 <span class="hlt">keV</span>), INTEGRAL ISGRI (20-150 <span class="hlt">keV</span>), and Fermi GBM (20-300 <span class="hlt">keV</span>). The pulsed flux (30-100 MeV) is (3.91 ± 0.97) × 10-9 photons cm-2 s-1 MeV-1. Declining significances of the INTEGRAL ISGRI 20-150 <span class="hlt">keV</span> pulse profiles suggest fading of the pulsed hard X-ray emission during the post-outburst epochs. We revisited with greatly improved statistics the timing and spectral characteristics of PSR B1509-58 as measured with the Fermi LAT. The broad-<span class="hlt">band</span> pulsed emission spectra (from 2 <span class="hlt">keV</span> up to GeV <span class="hlt">energies</span>) of PSR J1846-0258 and PSR B1509-58 can be accurately described with similarly curved shapes, with maximum luminosities at 3.5 ± 1.1 MeV (PSR J1846-0258) and 2.23 ± 0.11 MeV (PSR B1509-58). We discuss possible explanations for observational differences between Fermi LAT detected pulsars that reach maximum luminosities at GeV <span class="hlt">energies</span>, like the second magnetar-like pulsar PSR J1119-6127, and pulsars with maximum luminosities at MeV <span class="hlt">energies</span>, which might be due to geometric differences rather than exotic physics in high-B fields.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017ApJ...846...20L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017ApJ...846...20L"><span>The NuSTAR Serendipitous Survey: Hunting for the Most Extreme Obscured AGN at >10 <span class="hlt">keV</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lansbury, G. B.; Alexander, D. M.; Aird, J.; Gandhi, P.; Stern, D.; Koss, M.; Lamperti, I.; Ajello, M.; Annuar, A.; Assef, R. J.; Ballantyne, D. R.; Baloković, M.; Bauer, F. E.; Brandt, W. N.; Brightman, M.; Chen, C.-T. J.; Civano, F.; Comastri, A.; Del Moro, A.; Fuentes, C.; Harrison, F. A.; Marchesi, S.; Masini, A.; Mullaney, J. R.; Ricci, C.; Saez, C.; Tomsick, J. A.; Treister, E.; Walton, D. J.; Zappacosta, L.</p> <p>2017-09-01</p> <p>We identify sources with extremely hard X-ray spectra (I.e., with photon indices of {{Γ }}≲ 0.6) in the 13 deg2 NuSTAR serendipitous survey, to search for the most highly obscured active galactic nuclei (AGNs) detected at > 10 {<span class="hlt">keV</span>}. Eight extreme NuSTAR sources are identified, and we use the NuSTAR data in combination with lower-<span class="hlt">energy</span> X-ray observations (from Chandra, Swift XRT, and XMM-Newton) to characterize the broadband (0.5-24 <span class="hlt">keV</span>) X-ray spectra. We find that all of the extreme sources are highly obscured AGNs, including three robust Compton-thick (CT; {N}{{H}}> 1.5× {10}24 cm-2) AGNs at low redshift (z< 0.1) and a likely CT AGN at higher redshift (z = 0.16). Most of the extreme sources would not have been identified as highly obscured based on the low-<span class="hlt">energy</span> (< 10 <span class="hlt">keV</span>) X-ray coverage alone. The multiwavelength properties (e.g., optical spectra and X-ray-mid-IR luminosity ratios) provide further support for the eight sources being significantly obscured. Correcting for absorption, the intrinsic rest-frame 10-40 <span class="hlt">keV</span> luminosities of the extreme sources cover a broad range, from ≈ 5× {10}42 to 1045 erg s-1. The estimated number counts of CT AGNs in the NuSTAR serendipitous survey are in broad agreement with model expectations based on previous X-ray surveys, except for the lowest redshifts (z< 0.07), where we measure a high CT fraction of {f}{CT}{obs}={30}-12+16 % . For the small sample of CT AGNs, we find a high fraction of galaxy major mergers (50% ± 33%) compared to control samples of “normal” AGNs.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JPhCS.940a2041F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JPhCS.940a2041F"><span>Measurement of the strengths of the resonances at 417, 458, 611, 632 and 1222 <span class="hlt">keV</span> in the 22Ne(p, γ)23Na reaction</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ferraro, Federico</p> <p>2018-01-01</p> <p>The 22Ne(p, γ)23Na reaction is part of the NeNa cycle of hydrogen burning. This cycle plays a key role in the nucleosynthesis of the elements between 20Ne and 27Al in red giant stars, asymptotic giant stars and classical nova explosions. The strengths of the resonances at proton <span class="hlt">energies</span> above 400 <span class="hlt">keV</span> are still affected by high uncertainty. In order to reduce this uncertainty, a precision study of some of the most intense resonances between 400 <span class="hlt">keV</span> and 1250 <span class="hlt">keV</span> has been performed at the HZDR 3 MV Tandetron. The target, made of 22Ne implanted in a 0.22 mm thick Ta backing, has been characterized using the 1222 <span class="hlt">keV</span> and 458 <span class="hlt">keV</span> resonances, well known in literature. Subsequently, the strengths of the resonances at 417, 458, 611, 632 and 1222 <span class="hlt">keV</span> were determined. Two HPGe detectors equipped with active anti-Compton shielding have been used.</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/2017PhRvC..96f4601W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017PhRvC..96f4601W"><span>Measurement of the 238U(n ,γ ) cross section up to 80 <span class="hlt">keV</span> with the Total Absorption Calorimeter at the CERN n_TOF facility</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wright, T.; Guerrero, C.; Billowes, J.; Cano-Ott, D.; Mendoza, E.; Altstadt, S.; Andrzejewski, J.; Audouin, L.; Bécares, V.; Barbagallo, M.; Bečvář, F.; Belloni, F.; Berthoumieux, E.; Bosnar, D.; Brugger, M.; Calviño, F.; Calviani, M.; Carrapiço, C.; Cerutti, F.; Chiaveri, E.; Chin, M.; Colonna, N.; Cortés, G.; Cortés-Giraldo, M. A.; Diakaki, M.; Dietz, M.; Domingo-Pardo, C.; Durán, I.; Dzysiuk, N.; Eleftheriadis, C.; Ferrari, A.; Fraval, K.; Furman, V.; Gómez-Hornillos, M. B.; Ganesan, S.; García, A. R.; Giubrone, G.; Gonçalves, I. F.; González-Romero, E.; Goverdovski, A.; Griesmayer, E.; Gunsing, F.; Gurusamy, P.; Heftrich, T.; Hernández-Prieto, A.; Jenkins, D. G.; Jericha, E.; Käppeler, F.; Kadi, Y.; Karadimos, D.; Katabuchi, T.; Ketlerov, V.; Khryachkov, V.; Koehler, P.; Kokkoris, M.; Kroll, J.; Krtička, M.; Lampoudis, C.; Langer, C.; Leal-Cidoncha, E.; Lederer, C.; Leeb, H.; Leong, L. S.; Lerendegui-Marco, J.; Losito, R.; Manousos, A.; Marganiec, J.; Martínez, T.; Massimi, C.; Mastinu, P.; Mengoni, A.; Milazzo, P. M.; Mingrone, F.; Mirea, M.; Paradela, C.; Pavlik, A.; Perkowski, J.; Praena, J.; Quesada, J. M.; Rauscher, T.; Reifarth, R.; Riego-Perez, A.; Robles, M.; Roman, F.; Rubbia, C.; Ryan, J. A.; Sabaté-Gilarte, M.; Sarmento, R.; Saxena, A.; Schillebeeckx, P.; Schmidt, S.; Schumann, D.; Sedyshev, P.; Tagliente, G.; Tain, J. L.; Tarifeño-Saldivia, A.; Tarrío, D.; Tassan-Got, L.; Tsinganis, A.; Valenta, S.; Vannini, G.; Variale, V.; Vaz, P.; Ventura, A.; Vermeulen, M. J.; Versaci, R.; Vlachoudis, V.; Vlastou, R.; Wallner, A.; Ware, T.; Weigand, M.; Weiss, C.; Žugec, P.; n TOF Collaboration</p> <p>2017-12-01</p> <p>The radiative capture cross section of a highly pure (99.999%), 6.125(2) grams and 9.56(5)×10-4 atoms/barn areal density 238U sample has been measured with the Total Absorption Calorimeter (TAC) in the 185 m flight path at the CERN neutron time-of-flight facility n_TOF. This measurement is in response to the NEA High Priority Request list, which demands an accuracy in this cross section of less than 3% below 25 <span class="hlt">keV</span>. These data have undergone careful background subtraction, with special care being given to the background originating from neutrons scattered by the 238U sample. Pileup and dead-time effects have been corrected for. The measured cross section covers an <span class="hlt">energy</span> range between 0.2 eV and 80 <span class="hlt">keV</span>, with an accuracy that varies with neutron <span class="hlt">energy</span>, being better than 4% below 25 <span class="hlt">keV</span> and reaching at most 6% at higher <span class="hlt">energies</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20130013642','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20130013642"><span>GRB110721A: An Extreme Peak <span class="hlt">Energy</span> and Signatures of the Photosphere</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Axelsson, M.; Baldini, L.; Barbiellini, G.; Baring, M. G.; Bellazzini, R.; Bregeon, J.; Brigida, M.; Bruel, P.; Buehler, R.; Caliandro, G. A.; <a style="text-decoration: none; " href="javascript:void(0); " onClick="displayelement('author_20130013642'); toggleEditAbsImage('author_20130013642_show'); toggleEditAbsImage('author_20130013642_hide'); "> <img style="display:inline; width:12px; height:12px; " src="images/arrow-up.gif" width="12" height="12" border="0" alt="hide" id="author_20130013642_show"> <img style="width:12px; height:12px; display:none; " src="images/arrow-down.gif" width="12" height="12" border="0" alt="hide" id="author_20130013642_hide"></p> <p>2012-01-01</p> <p>GRB110721A was observed by the Fermi Gamma-ray Space Telescope using its two instruments, the Large Area Telescope (LAT) and the Gamma-ray Burst Monitor (GBM). The burst consisted of one major emission episode which lasted for approximately 24.5 s (in the GBM) and had a peak flux of (5.7 +/- 0.2) 10(exp -5) erg s(exp -1) cm(exp -2). The time-resolved emission spectrum is best modeled with a combination of a <span class="hlt">Band</span> function and a blackbody spectrum. The peak <span class="hlt">energy</span> of the <span class="hlt">Band</span> component was initially 15 +/- 2 MeV, which is the highest value ever detected in a GRB. This measurement was made possible by combining GBM/BGO data with LAT Low <span class="hlt">Energy</span> events to achieve continuous 10-100 MeV coverage. The peak <span class="hlt">energy</span> later decreased as a power law in time with an index of -1.89 +/- 0.10. The temperature of the blackbody component also decreased, starting from approximately 80 <span class="hlt">keV</span>, and the decay showed a significant break after approximately 2s. The spectrum provides strong constraints on the standard synchrotron model, indicating that alternative mechanisms may give rise to the emission at these <span class="hlt">energies</span>.</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 additional 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('https://www.osti.gov/biblio/22518583-quiet-time-suprathermal-kev-electrons-solar-wind','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22518583-quiet-time-suprathermal-kev-electrons-solar-wind"><span>QUIET-TIME SUPRATHERMAL (∼0.1–1.5 <span class="hlt">keV</span>) ELECTRONS IN THE SOLAR WIND</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>Tao, Jiawei; Wang, Linghua; Zong, Qiugang</p> <p>2016-03-20</p> <p>We present a statistical survey of the <span class="hlt">energy</span> spectrum of solar wind suprathermal (∼0.1–1.5 <span class="hlt">keV</span>) electrons measured by the WIND 3DP instrument at 1 AU during quiet times at the minimum and maximum of solar cycles 23 and 24. After separating (beaming) strahl electrons from (isotropic) halo electrons according to their different behaviors in the angular distribution, we fit the observed <span class="hlt">energy</span> spectrum of both strahl and halo electrons at ∼0.1–1.5 <span class="hlt">keV</span> to a Kappa distribution function with an index κ and effective temperature T{sub eff}. We also calculate the number density n and average <span class="hlt">energy</span> E{sub avg} of strahl andmore » halo electrons by integrating the electron measurements between ∼0.1 and 1.5 <span class="hlt">keV</span>. We find a strong positive correlation between κ and T{sub eff} for both strahl and halo electrons, and a strong positive correlation between the strahl n and halo n, likely reflecting the nature of the generation of these suprathermal electrons. In both solar cycles, κ is larger at solar minimum than at solar maximum for both strahl and halo electrons. The halo κ is generally smaller than the strahl κ (except during the solar minimum of cycle 23). The strahl n is larger at solar maximum, but the halo n shows no difference between solar minimum and maximum. Both the strahl n and halo n have no clear association with the solar wind core population, but the density ratio between the strahl and halo roughly anti-correlates (correlates) with the solar wind density (velocity)« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20020050368&hterms=sources+energy&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dsources%2Benergy','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20020050368&hterms=sources+energy&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dsources%2Benergy"><span>The BATSE Earth Occultation Catalog of Low <span class="hlt">Energy</span> Gamma Ray Sources</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Harmon, B. A.; Wilson-Hodge, C. A.; Fishman, G. J.; Paciesas, W. S.; Zhang, S. N.; Finger, M. H.; Connaughton, V.; Koshut, T. M.; Henze, W.; McCollough, M. L.; <a style="text-decoration: none; " href="javascript:void(0); " onClick="displayelement('author_20020050368'); toggleEditAbsImage('author_20020050368_show'); toggleEditAbsImage('author_20020050368_hide'); "> <img style="display:inline; width:12px; height:12px; " src="images/arrow-up.gif" width="12" height="12" border="0" alt="hide" id="author_20020050368_show"> <img style="width:12px; height:12px; display:none; " src="images/arrow-down.gif" width="12" height="12" border="0" alt="hide" id="author_20020050368_hide"></p> <p>2002-01-01</p> <p>The Burst and Transient Source Experiment (BATSE), aboard the Compton Gamma Ray Observatory (CGRO), provided a record of the hard X-ray/low <span class="hlt">energy</span> gamma ray sky between April 1991 and June 2000. During that time, a catalog of known sources was derived from existing catalogs such as HEAO A-4 (Levine et al. 1984), as well as new transient sources discovered with BATSE and other X-ray monitors operating in the CGRO era. The Earth Occultation Technique (Harmon et al. 2001, astro-ph/0109069) was used to monitor a combination of these sources, mostly galactic, totaling to about 175 objects. The catalog will present the global properties of these sources and their probability of detection (greater than 10 mCrab, 20-100 <span class="hlt">keV</span>) with BATSE. Systematic errors due to unknown sources or background components are included. Cursory analyses to search for new transients (35-80 mCrab in the 20-100 <span class="hlt">keV</span> <span class="hlt">band</span>) and super-orbital periods in known binary sources are also presented. Whole mission light curves and associated data production/analysis tools are being delivered to the High <span class="hlt">Energy</span> Astrophysics Science Archive Research Center (HEASARC) for public use.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014EPJA...50...55W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014EPJA...50...55W"><span>Study on the <span class="hlt">keV</span> neutron capture reaction in 56Fe and 57Fe</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wang, Taofeng; Lee, Manwoo; Kim, Guinyun; Ro, Tae-Ik; Kang, Yeong-Rok; Igashira, Masayuki; Katabuchi, Tatsuya</p> <p>2014-03-01</p> <p>The neutron capture cross-sections and the radiative capture gamma-ray spectra from the broad resonances of 56Fe and 57Fe in the neutron <span class="hlt">energy</span> range from 10 to 90<span class="hlt">keV</span> and 550<span class="hlt">keV</span> have been measured with an anti-Compton NaI(Tl) detector. Pulsed <span class="hlt">keV</span> neutrons were produced from the 7Li 7Be reaction by bombarding the lithium target with the 1.5ns bunched proton beam from the 3MV Pelletron accelerator. The incident neutron spectrum on a capture sample was measured by means of a time-of-flight (TOF) method with a 6Li -glass detector. The number of weighted capture counts of the iron or gold sample was obtained by applying a pulse height weighting technique to the corresponding capture gamma-ray pulse height spectrum. The neutron capture gamma-ray spectra were obtained by unfolding the observed capture gamma-ray pulse height spectra. To achieve further understanding on the mechanism of neutron radiative capture reaction and study on physics models, theoretical calculations of the -ray spectra for 56Fe and 57Fe with the POD program have been performed by applying the Hauser-Feshbach statistical model. The dominant ingredients to perform the statistical calculation were the Optical Model Potential (OMP), the level densities described by the Mengoni-Nakajima approach, and the -ray transmission coefficients described by -ray strength functions. The comparison of the theoretical calculations, performed only for the 550<span class="hlt">keV</span> point, show a good agreement with the present experimental results.</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 addition, 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 addition, 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/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> </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('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/2018PhRvL.120f1301B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PhRvL.120f1301B"><span>X-Ray Lines from Dark Matter Annihilation at the <span class="hlt">keV</span> Scale</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Brdar, Vedran; Kopp, Joachim; Liu, Jia; Wang, Xiao-Ping</p> <p>2018-02-01</p> <p>In 2014, several groups reported hints for a yet unidentified line in astrophysical x-ray signals from galaxies and galaxy clusters at an <span class="hlt">energy</span> of 3.5 <span class="hlt">keV</span>. While it is not unlikely that this line is simply a reflection of imperfectly modeled atomic transitions, it has renewed the community's interest in models of keV-scale dark matter, whose decay would lead to such a line. The alternative possibility of dark matter annihilation into monochromatic photons is far less explored, a lapse that we strive to amend in this Letter. More precisely, we introduce a novel model of fermionic dark matter χ with O (<span class="hlt">keV</span> ) mass, annihilating to a scalar state ϕ which in turn decays to photons, for instance via loops of heavy vectorlike fermions. The resulting photon spectrum is box shaped, but if χ and ϕ are nearly degenerate in mass, it can also resemble a narrow line. We discuss dark matter production via two different mechanisms—misalignment and freeze-in—which both turn out to be viable in vast regions of parameter space. We constrain the model using astrophysical x-ray data, and we demonstrate that, thanks to the velocity dependence of the annihilation cross section, it has the potential to reconcile the various observations of the 3.5 <span class="hlt">keV</span> line. We finally argue that the model can easily avoid structure formation constraints on keV-scale dark matter.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1402656-generation-lower-upper-bands-electrostatic-electron-cyclotron-harmonic-waves-van-allen-radiation-belts','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1402656-generation-lower-upper-bands-electrostatic-electron-cyclotron-harmonic-waves-van-allen-radiation-belts"><span>Generation of lower and upper <span class="hlt">bands</span> of electrostatic electron cyclotron harmonic waves in the Van Allen radiation belts</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Zhou, Qinghua; Xiao, Fuliang; Yang, Chang; ...</p> <p>2017-05-22</p> <p>Electrostatic electron cyclotron harmonic (ECH) waves generated by the electron loss cone distribution can produce efficient scattering loss of plasma sheet electrons, which has a significant effect on the dynamics in the outer magnetosphere. Here we report two ECH emission events around the same location L≈ 5.7–5.8, MLT ≈ 12 from Van Allen Probes on 11 February (event A) and 9 January 2014 (event B), respectively. The spectrum of ECH waves was centered at the lower half of the harmonic <span class="hlt">bands</span> during event A, but the upper half during event B. The observed electron phase space density in both eventsmore » is fitted by the subtracted bi-Maxwellian distribution, and the fitting functions are used to evaluate the local growth rates of ECH waves based on a linear theory for homogeneous plasmas. ECH waves are excited by the loss cone instability of 50 eV–1 <span class="hlt">keV</span> electrons in the lower half of harmonic <span class="hlt">bands</span> in the low-density plasmasphere in event A, and 1–10 <span class="hlt">keV</span> electrons in the upper half of harmonic <span class="hlt">bands</span> in a relatively high-density region in event B. Here, the current results successfully explain observations and provide a first direct evidence on how ECH waves are generated in the lower and upper half of harmonic frequency <span class="hlt">bands</span>.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1402656','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1402656"><span>Generation of lower and upper <span class="hlt">bands</span> of electrostatic electron cyclotron harmonic waves in the Van Allen radiation belts</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>Zhou, Qinghua; Xiao, Fuliang; Yang, Chang</p> <p></p> <p>Electrostatic electron cyclotron harmonic (ECH) waves generated by the electron loss cone distribution can produce efficient scattering loss of plasma sheet electrons, which has a significant effect on the dynamics in the outer magnetosphere. Here we report two ECH emission events around the same location L≈ 5.7–5.8, MLT ≈ 12 from Van Allen Probes on 11 February (event A) and 9 January 2014 (event B), respectively. The spectrum of ECH waves was centered at the lower half of the harmonic <span class="hlt">bands</span> during event A, but the upper half during event B. The observed electron phase space density in both eventsmore » is fitted by the subtracted bi-Maxwellian distribution, and the fitting functions are used to evaluate the local growth rates of ECH waves based on a linear theory for homogeneous plasmas. ECH waves are excited by the loss cone instability of 50 eV–1 <span class="hlt">keV</span> electrons in the lower half of harmonic <span class="hlt">bands</span> in the low-density plasmasphere in event A, and 1–10 <span class="hlt">keV</span> electrons in the upper half of harmonic <span class="hlt">bands</span> in a relatively high-density region in event B. Here, the current results successfully explain observations and provide a first direct evidence on how ECH waves are generated in the lower and upper half of harmonic frequency <span class="hlt">bands</span>.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19980111017','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19980111017"><span>Atomic, Molecular, and Optical Physics: Optical Excitation Function of H(1s-2p) Produced by electron Impact from Threshold to 1.8 <span class="hlt">keV</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>James, G. K.; Slevin, J. A.; Shemansky, D. E.; McConkey, J. W.; Bray, I.; Dziczek, D.; Kanik, I.; Ajello, J. M.</p> <p>1997-01-01</p> <p>The optical excitation function of prompt Lyman-Alpha radiation, produced by electron impact on atomic hydrogen, has been measured over the extended <span class="hlt">energy</span> range from threshold to 1.8 <span class="hlt">keV</span>. Measurements were obtained in a crossed-beams experiment using both magnetically confined and electrostatically focused electrons in collision with atomic hydrogen produced by an intense discharge source. A vacuum-ultraviolet mono- chromator system was used to measure the emitted Lyman-Alpha radiation. The absolute H(1s-2p) electron impact excitation cross section was obtained from the experimental optical excitation function by normalizing to the accepted optical oscillator strength, with corrections for polarization and cascade. Statistical and known systematic uncertainties in our data range from +/- 4% near threshold to +/- 2% at 1.8 <span class="hlt">keV</span>. Multistate coupling affecting the shape of the excitation function up to 1 <span class="hlt">keV</span> impact <span class="hlt">energy</span> is apparent in both the present experimental data and present theoretical results obtained with convergent close- coupling (CCC) theory. This shape function effect leads to an uncertainty in absolute cross sections at the 10% level in the analysis of the experimental data. The derived optimized absolute cross sections are within 7% of the CCC calculations over the 14 eV-1.8 <span class="hlt">keV</span> range. The present CCC calculations converge on the Bethe- Fano profile for H(1s-2p) excitation at high <span class="hlt">energy</span>. For this reason agreement with the CCC values to within 3% is achieved in a nonoptimal normalization of the experimental data to the Bethe-Fano profile. The fundamental H(1s-2p) electron impact cross section is thereby determined to an unprecedented accuracy over the 14 eV - 1.8 <span class="hlt">keV</span> <span class="hlt">energy</span> range.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19740049184&hterms=centaurus&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D90%26Ntt%3Dcentaurus','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19740049184&hterms=centaurus&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D90%26Ntt%3Dcentaurus"><span>Extended observations of higher than 7-<span class="hlt">keV</span> X-rays from Centaurus X-3 by the OSO-7 satellite</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Baity, W. A.; Ulmer, M. P.; Wheaton, W. A.; Peterson, L. E.</p> <p>1974-01-01</p> <p>The UCSD X-ray telescope on board OSO-7 provided 43 days of continuous coverage of the variable X-ray source Cen X-3 at <span class="hlt">energies</span> above 7 <span class="hlt">keV</span> during December 1971 and January 1972. We detected the 4.8-sec pulsation period, the 2.087-day eclipse cycle, and an apparently nonperiodic, low-intensity state lasting more than 12 days. Spectra obtained over the 7-30 <span class="hlt">keV</span> range during noneclipsed high-intensity states are steeper than those previously reported. Large changes, which may be characterized by a number spectral index alpha varying between 3.0 plus or minus 0.2 and 2.0 plus or minus 0.3, or by exponential spectra with kT varying from 6 plus or minus 2 to 13 plus or minus 3 <span class="hlt">keV</span>, occur at different high-intensity states.</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/2010APS..DNP.GE002H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010APS..DNP.GE002H"><span>High-precision measurement of the light response of BC-418 plastic scintillator to protons with <span class="hlt">energies</span> from 100 <span class="hlt">keV</span> to 10 MeV</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Henzl, Vladimir; Daub, Brian; French, Jennifer; Matthews, June; Kovash, Michael; Wender, Stephen; Famiano, Michael; Koehler, Katrina; Yuly, Mark</p> <p>2010-11-01</p> <p>The determination of the light response of many organic scintillators to various types of radiation has been a subject of numerous experimental as well as theoretical studies in the past. But while the data on light response to particles with <span class="hlt">energies</span> above 1 MeV are precise and abundant, the information on light response to very low <span class="hlt">energy</span> particles (i.e. below 1 MeV) is scarce or completely missing. In this study we measured the light response of a BC-418 scintillator to protons with <span class="hlt">energies</span> from 100 <span class="hlt">keV</span> to 10 MeV. The experiment was performed at Weapons Neutron Research Facility at LANSCE, Los Alamos. The neutron beam from a spallation source is used to irradiate the active target made from BC-418 plastic scintillator. The recoiled protons detected in the active target are measured in coincidence with elastically scattered incident neutrons detected by and adjacent liquid scintillator. Time of flight of the incident neutron and the knowledge of scattering geometry allow for a kinematically complete and high-precision measurement of the light response as a function of the proton <span class="hlt">energy</span>.</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('https://www.ncbi.nlm.nih.gov/pubmed/9232893','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/9232893"><span>Induction of single- and double-strand breaks in plasmid DNA by monoenergetic alpha-particles with <span class="hlt">energies</span> below the Bragg-maximum.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Scholz, V; Weidner, J; Köhnlein, W; Frekers, D; Wörtche, H J</p> <p>1997-01-01</p> <p>The yield of single-strand breaks (ssb) and double-strand breaks (dsb) produced by alpha-particles at the end of their track in DNA-films was determined experimentally. Helium nuclei were accelerated to 600 <span class="hlt">keV</span> in the 400 kV ion accelerator and scattered at a carbon target. The elastically scattered alpha-particles with <span class="hlt">energies</span> of 344 <span class="hlt">keV</span> and 485 <span class="hlt">keV</span> were used to irradiate supercircular plasmid DNA in vacuo. For the dosimetry of the alpha-particles a surface barrier detector was used and the <span class="hlt">energy</span> distribution of the alpha-particles determined. The <span class="hlt">energy</span> loss of the particles in the DNA-layer was calculated. DNA samples were separated into the three conformational isomers using agarose gel electrophoresis. After fluorochromation the number of ssb and dsb per plasmid DNA molecule was established from the <span class="hlt">band</span> intensities assuming the validity of Poisson statistics. Linear dose effect correlations were found for ssb and dsb per plasmid molecule. In the case of 344 keV-alpha-particles the yield of dsb was (8.6 +/- 0.9) x 10(-11) breaks/Gy x dalton. The ratio of ssb/dsb was 0.5 +/- 0.2. This is at least a factor of six larger than the ratio found in experiments with higher <span class="hlt">energy</span> alpha-particles and from model calculations. Similar experiments with protons yielded a relative biological effectiveness (rbe) value of 2.8 for the induction of double-strand breaks by track end alpha-particles.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22522288-stability-extraterrestrial-glycine-under-energetic-particle-radiation-estimated-from-kev-electron-bombardment-experiments','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22522288-stability-extraterrestrial-glycine-under-energetic-particle-radiation-estimated-from-kev-electron-bombardment-experiments"><span>STABILITY OF EXTRATERRESTRIAL GLYCINE UNDER ENERGETIC PARTICLE RADIATION ESTIMATED FROM 2 <span class="hlt">keV</span> ELECTRON BOMBARDMENT EXPERIMENTS</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, B.; Tanarro, I.; Escribano, R.</p> <p>2015-06-20</p> <p>The destruction of solid glycine under irradiation with 2 <span class="hlt">keV</span> electrons has been investigated by means of IR spectroscopy. Destruction cross sections, radiolysis yields, and half-life doses were determined for samples at 20, 40, 90, and 300 K. The thickness of the irradiated samples was kept below the estimated penetration depth of the electrons. No significant differences were obtained in the experiments below 90 K, but the destruction cross section at 300 K was larger by a factor of 2. The radiolysis yields and half-life doses are in good accordance with recent MeV proton experiments, which confirms that electrons inmore » the <span class="hlt">keV</span> range can be used to simulate the effects of cosmic rays if the whole sample is effectively irradiated. In the low temperature experiments, electron irradiation leads to the formation of residues. IR absorptions of these residues are assigned to the presence CO{sub 2}, CO, OCN{sup −}, and CN{sup −} and possibly to amide <span class="hlt">bands</span> I to III. The protection of glycine by water ice is also studied. A water ice film of ∼150 nm is found to provide efficient shielding against the bombardment of 2 <span class="hlt">keV</span> electrons. The results of this study show also that current Monte Carlo predictions provide a good global description of electron penetration depths. The lifetimes estimated in this work for various environments ranging from the diffuse interstellar medium to the inner solar system, show that the survival of hypothetical primeval glycine from the solar nebula in present solar system bodies is not very likely.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018MNRAS.473.4360W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018MNRAS.473.4360W"><span>Super-Eddington accretion on to the neutron star NGC 7793 P13: Broad-<span class="hlt">band</span> X-ray spectroscopy and ultraluminous X-ray sources</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Walton, D. J.; Fürst, F.; Harrison, F. A.; Stern, D.; Bachetti, M.; Barret, D.; Brightman, M.; Fabian, A. C.; Middleton, M. J.; Ptak, A.; Tao, L.</p> <p>2018-02-01</p> <p>We present a detailed, broad-<span class="hlt">band</span> X-ray spectral analysis of the ultraluminous X-ray source (ULX) pulsar NGC 7793 P13, a known super-Eddington source, utilizing data from the XMM-Newton, NuSTAR and Chandra observatories. The broad-<span class="hlt">band</span> XMM-Newton+NuSTAR spectrum of P13 is qualitatively similar to the rest of the ULX sample with broad-<span class="hlt">band</span> coverage, suggesting that additional ULXs in the known population may host neutron star accretors. Through time-averaged, phase-resolved and multi-epoch studies, we find that two non-pulsed thermal blackbody components with temperatures ∼0.5 and 1.5 <span class="hlt">keV</span> are required to fit the data below 10 <span class="hlt">keV</span>, in addition to a third continuum component which extends to higher <span class="hlt">energies</span> and is associated with the pulsed emission from the accretion column. The characteristic radii of the thermal components appear to be comparable, and are too large to be associated with the neutron star itself, so the need for two components likely indicates the accretion flow outside the magnetosphere is complex. We suggest a scenario in which the thick inner disc expected for super-Eddington accretion begins to form, but is terminated by the neutron star's magnetic field soon after its onset, implying a limit of B ≲ 6 × 1012 G for the dipolar component of the central neutron star's magnetic field. Evidence of similar termination of the disc in other sources may offer a further means of identifying additional neutron star ULXs. Finally, we examine the spectrum exhibited by P13 during one of its unusual 'off' states. These data require both a hard power-law component, suggesting residual accretion on to the neutron star, and emission from a thermal plasma, which we argue is likely associated with the P13 system.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JPhB...50u5002L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JPhB...50u5002L"><span>Study of inelastic processes in Li+-Ar, K+-Ar, and Na+-He collisions in the <span class="hlt">energy</span> range 0.5-10 <span class="hlt">keV</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lomsadze, Ramaz A.; Gochitashvili, Malkhaz R.; Kezerashvili, Roman Ya; Schulz, Michael</p> <p>2017-11-01</p> <p>Absolute cross sections are measured for charge-exchange, ionization, and excitation processes within the same experimental setup for the Li{}+-Ar, K{}+-Ar, and Na{}+-He collisions in the ion <span class="hlt">energy</span> range of 0.5-10 <span class="hlt">keV</span>. The results of the measurements and schematic correlation diagrams are used to analyze and determine the mechanisms for these processes. The experimental results show that the charge-exchange processes occur with high probabilities and electrons are predominantly captured in ground states. The contributions of various partial inelastic channels to the total ionization cross section are estimated, and a primary mechanism for the process is identified. In addition, the <span class="hlt">energy</span>-loss spectrum is applied in order to estimate the relative contribution of different inelastic channels, and to determine the mechanisms for the ionization and for some excitation processes of Ar resonance lines for the {{{K}}}+-Ar collision system. The excitation cross sections for the helium and for the sodium doublet lines for the Na{}+-He collision system both reveal some unexpected features. A mechanism to explain this observation is suggested.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018RaPC..148...73O','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018RaPC..148...73O"><span>Performance of different theories for the angular distribution of bremsstrahlung produced by <span class="hlt">keV</span> electrons incident upon a target</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Omar, Artur; Andreo, Pedro; Poludniowski, Gavin</p> <p>2018-07-01</p> <p>Different theories of the intrinsic bremsstrahlung angular distribution (i.e., the shape function) have been evaluated using Monte Carlo calculations for various target materials and incident electron <span class="hlt">energies</span> between 20 <span class="hlt">keV</span> and 300 <span class="hlt">keV</span>. The shape functions considered were the plane-wave first Born approximation cross sections (i) 2BS [high-<span class="hlt">energy</span> result, screened nucleus], (ii) 2BN [general result, bare nucleus], (iii) KM [2BS modified to emulate 2BN], and (iv) SIM [leading term of 2BN]; (v) expression based on partial-waves expansion, KQP; and (vi) a uniform spherical distribution, UNI [a common approximation in certain analytical models]. The shape function was found to have an important impact on the bremsstrahlung emerging from thin foil targets in which the incident electrons undergo few elastic scatterings before exiting the target material. For thick transmission and reflection targets the type of shape function had less importance, as the intrinsic bremsstrahlung angular distribution was masked by the diffuse directional distribution of multiple scattered electrons. Predictions made using the 2BN and KQP theories were generally in good agreement, suggesting that the effect of screening and the constraints of the Born approximation on the intrinsic angular distribution may be acceptable. The KM and SIM shape functions deviated notably from KQP for low electron <span class="hlt">energies</span> (< 50 <span class="hlt">keV</span>), while 2BS and UNI performed poorly over most of the <span class="hlt">energy</span> range considered; the 2BS shape function was found to be too forward-focused in emission, while UNI was not forward-focused enough. The results obtained emphasize the importance of the intrinsic bremsstrahlung angular distribution for theoretical predictions of x-ray emission, which is relevant in various applied disciplines, including x-ray crystallography, electron-probe microanalysis, security and industrial inspection, medical imaging, as well as low- and medium (orthovoltage) <span class="hlt">energy</span> radiotherapy.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19940035082&hterms=pietsch&qs=N%3D0%26Ntk%3DAuthor-Name%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dpietsch','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19940035082&hterms=pietsch&qs=N%3D0%26Ntk%3DAuthor-Name%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dpietsch"><span>Spectral and temporal properties of the X-ray pulsar SMC X-1 at hard X-rays</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Kunz, M.; Gruber, D. E.; Kendziorra, E .; Kretschmar, P.; Maisack, M.; Mony, B.; Staubert, R.; Doebereiner, S.; Englhauser, J.; Pietsch, W.</p> <p>1993-01-01</p> <p>The binary X-ray pulsar SMC X- 1 has been observed at hard X-rays with the High <span class="hlt">Energy</span> X-Ray Experiment (HEXE) on nine occasions between Nov. 1987 and March 1989. A thin thermal bremsstrahlung fit to the phase averaged spectrum yields a plasma temperature (14.4 +/- 1.3) <span class="hlt">keV</span> and a luminosity above (1.1 +/- 0.1) x 10 exp 38 erg/s in the 20-80 <span class="hlt">keV</span> <span class="hlt">band</span>. Pulse period values have been established for three observations, confirming the remarkably stable spin-up trend of SMC X-1. In one of the three observations the pulse profile was seen to deviate from a dominant double pulsation, while at the same time the pulsed fraction was unusually large. For one observation we determined for the first time the pulsed fraction in narrow <span class="hlt">energy</span> <span class="hlt">bands</span>. It increases with photon <span class="hlt">energy</span> from about 20 percent up to over 60 percent in the <span class="hlt">energy</span> range from 20 to 80 <span class="hlt">keV</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22522039-broadband-ray-imaging-spectroscopy-crab-nebula-pulsar-nustar','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22522039-broadband-ray-imaging-spectroscopy-crab-nebula-pulsar-nustar"><span>BROADBAND X-RAY IMAGING AND SPECTROSCOPY OF THE CRAB NEBULA AND PULSAR WITH NuSTAR</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>Madsen, Kristin K.; Harrison, Fiona; Grefenstette, Brian W.</p> <p></p> <p>We present broadband (3-78 <span class="hlt">keV</span>) NuSTAR X-ray imaging and spectroscopy of the Crab nebula and pulsar. We show that while the phase-averaged and spatially integrated nebula + pulsar spectrum is a power law in this <span class="hlt">energy</span> <span class="hlt">band</span>, spatially resolved spectroscopy of the nebula finds a break at ∼9 <span class="hlt">keV</span> in the spectral photon index of the torus structure with a steepening characterized by ΔΓ ∼ 0.25. We also confirm a previously reported steepening in the pulsed spectrum, and quantify it with a broken power law with break <span class="hlt">energy</span> at ∼12 <span class="hlt">keV</span> and ΔΓ ∼ 0.27. We present spectral maps ofmore » the inner 100'' of the remnant and measure the size of the nebula as a function of <span class="hlt">energy</span> in seven <span class="hlt">bands</span>. These results find that the rate of shrinkage with <span class="hlt">energy</span> of the torus size can be fitted by a power law with an index of γ = 0.094 ± 0.018, consistent with the predictions of Kennel and Coroniti. The change in size is more rapid in the NW direction, coinciding with the counter-jet where we find the index to be a factor of two larger. NuSTAR observed the Crab during the latter part of a γ-ray flare, but found no increase in flux in the 3-78 <span class="hlt">keV</span> <span class="hlt">energy</span> <span class="hlt">band</span>.« less</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 addition 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('https://ntrs.nasa.gov/search.jsp?R=19810046652&hterms=black+knight&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3Dblack%2Bknight','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19810046652&hterms=black+knight&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3Dblack%2Bknight"><span>Centaurus A /NGC 5128/ at 2 <span class="hlt">keV</span>-2.3 MeV - HEAO 1 observations and implications</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Baity, W. A.; Rothschild, R. E.; Lingenfelter, R. E.; Stein, W. A.; Nolan, P. L.; Gruber, D. E.; Knight, F. K.; Matteson, J. L.; Peterson, L. E.; Mushotzky, R. F.</p> <p>1981-01-01</p> <p>The active-nucleus galaxy Centaurus A has been studied at 2 <span class="hlt">keV</span>-2.3 MeV using data from the UCSD/MIT hard X-ray and low-<span class="hlt">energy</span> gamma-ray instrument and the GSFC/CIT cosmic X-ray experiment on HEAO-1. It is found that an E exp -1.60 + or - 0.03 power law spectrum breaking to E exp -2.0 + or - 0.2 at 140 <span class="hlt">keV</span> best describes the January and July 1978 data. The average intensity was 50% higher during the January observations. Upper limits to unresolved lines at 511 <span class="hlt">keV</span> and 1.6 MeV were found to be 6.5 x 10 to the -4th photons/sq cm-s and 2.2 x 10 to the -4th photons/sq cm-s, respectively, at the 90% confidence level. The present data are consistent with the detailed calculations of the synchrotron self-Compton mechanism; they may also agree, marginally, with the predictions of emission from spherical accretion onto black holes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1402625-characteristic-energy-range-electron-scattering-due-plasmaspheric-hiss','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1402625-characteristic-energy-range-electron-scattering-due-plasmaspheric-hiss"><span>Characteristic <span class="hlt">energy</span> range of electron scattering due to plasmaspheric hiss</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Ma, Q.; Li, W.; Thorne, R. M.; ...</p> <p>2016-11-15</p> <p>In this paper, we investigate the characteristic <span class="hlt">energy</span> range of electron flux decay due to the interaction with plasmaspheric hiss in the Earth's inner magnetosphere. The Van Allen Probes have measured the energetic electron flux decay profiles in the Earth's outer radiation belt during a quiet period following the geomagnetic storm that occurred on 7 November 2015. The observed <span class="hlt">energy</span> of significant electron decay increases with decreasing L shell and is well correlated with the <span class="hlt">energy</span> <span class="hlt">band</span> corresponding to the first adiabatic invariant μ = 4–200 MeV/G. The electron diffusion coefficients due to hiss scattering are calculated at L =more » 2–6, and the modeled <span class="hlt">energy</span> <span class="hlt">band</span> of effective pitch angle scattering is also well correlated with the constant μ lines and is consistent with the observed <span class="hlt">energy</span> range of electron decay. Using the previously developed statistical plasmaspheric hiss model during modestly disturbed periods, we perform a 2-D Fokker-Planck simulation of the electron phase space density evolution at L = 3.5 and demonstrate that plasmaspheric hiss causes the significant decay of 100 keV–1 MeV electrons with the largest decay rate occurring at around 340 <span class="hlt">keV</span>, forming anisotropic pitch angle distributions at lower <span class="hlt">energies</span> and more flattened distributions at higher <span class="hlt">energies</span>. Finally, our study provides reasonable estimates of the electron populations that can be most significantly affected by plasmaspheric hiss and the consequent electron decay profiles.« 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/10757312','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/10757312"><span>Oncogenic transformation in C3H10T1/2 cells by low-<span class="hlt">energy</span> neutrons.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Miller, R C; Marino, S A; Napoli, J; Shah, H; Hall, E J; Geard, C R; Brenner, D J</p> <p>2000-03-01</p> <p>Occupational exposure to neutrons typically includes significant doses of low-<span class="hlt">energy</span> neutrons, with <span class="hlt">energies</span> below 100 <span class="hlt">keV</span>. In addition, the normal-tissue dose from boron neutron capture therapy will largely be from low-<span class="hlt">energy</span> neutrons. Microdosimetric theory predicts decreasing biological effectiveness for neutrons with <span class="hlt">energies</span> below about 350 <span class="hlt">keV</span> compared with that for higher-<span class="hlt">energy</span> neutrons; based on such considerations, and limited biological data, the current radiation weighting factor (quality factor) for neutrons with <span class="hlt">energies</span> from 10 <span class="hlt">keV</span> to 100 <span class="hlt">keV</span> is less than that for higher-<span class="hlt">energy</span> neutrons. By contrast, some reports have suggested that the biological effectiveness of low-<span class="hlt">energy</span> neutrons is similar to that of fast neutrons. The purpose of the current work is to assess the relative biological effectiveness of low-<span class="hlt">energy</span> neutrons for an endpoint of relevance to carcinogenesis: in vitro oncogenic transformation. Oncogenic transformation induction frequencies were determined for C3H10T1/2 cells exposed to two low-<span class="hlt">energy</span> neutron beams, respectively, with dose-averaged <span class="hlt">energies</span> of 40 and 70 <span class="hlt">keV</span>, and the results were compared with those for higher-<span class="hlt">energy</span> neutrons and X-rays. These results for oncogenic transformation provide evidence for a significant decrease in biological effectiveness for 40 <span class="hlt">keV</span> neutrons compared with 350 <span class="hlt">keV</span> neutrons. The 70 <span class="hlt">keV</span> neutrons were intermediate in effectiveness between the 70 and 350 <span class="hlt">keV</span> beams. A decrease in biological effectiveness for low-<span class="hlt">energy</span> neutrons is in agreement with most (but not all) earlier biological studies, as well as microdosimetric considerations. The results for oncogenic transformation were consistent with the currently recommended decreased values for low-<span class="hlt">energy</span> neutron radiation weighting factors compared with fast neutrons.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/945646','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/945646"><span>Active detection of shielded SNM with 60-<span class="hlt">keV</span> neutrons</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>Hagmann, C; Dietrich, D; Hall, J</p> <p>2008-07-08</p> <p>Fissile materials, e.g. {sup 235}U and {sup 239}Pu, can be detected non-invasively by active neutron interrogation. A unique characteristic of fissile material exposed to neutrons is the prompt emission of high-<span class="hlt">energy</span> (fast) fission neutrons. One promising mode of operation subjects the object to a beam of medium-<span class="hlt">energy</span> (epithermal) neutrons, generated by a proton beam impinging on a Li target. The emergence of fast secondary neutrons then clearly indicates the presence of fissile material. Our interrogation system comprises a low-dose 60-<span class="hlt">keV</span> neutron generator (5 x 10{sup 6}/s), and a 1 m{sup 2} array of scintillators for fast neutron detection. Preliminary experimentalmore » results demonstrate the detectability of small quantities (370 g) of HEU shielded by steel (200 g/cm{sup 2}) or plywood (30 g/cm{sup 2}), with a typical measurement time of 1 min.« less</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/2015EPJD...69..235C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015EPJD...69..235C"><span>A coincidence study of electron and positron impact ionization of Ar (3p) at 1 <span class="hlt">keV</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Campeanu, Radu I.; Walters, James H. R.; Whelan, Colm T.</p> <p>2015-10-01</p> <p>Distorted-wave calculations of the triple differential cross section (TDCS) are presented for electron and positron impact ionization of Ar(3p) in coplanar asymmetric geometry at an impact <span class="hlt">energy</span> of 1 <span class="hlt">keV</span> and are compared with a recent experiment. The experiment indicates that the positron TDCS is generally larger than the equivalent electron TDCS. It is shown that the magnitude of the TDCS is extremely sensitive to the <span class="hlt">energy</span> of the ejected electron and that only when the cross section is averaged over <span class="hlt">energy</span> do we get a reasonable agreement with experiment.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2004AIPC..737..684D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2004AIPC..737..684D"><span>X-<span class="hlt">band</span> RF gun and linac for medical Compton scattering X-ray source</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Dobashi, Katsuhito; Uesaka, Mitsuru; Fukasawa, Atsushi; Sakamoto, Fumito; Ebina, Futaro; Ogino, Haruyuki; Urakawa, Junji; Higo, Toshiyasu; Akemoto, Mitsuo; Hayano, Hitoshi; Nakagawa, Keiichi</p> <p>2004-12-01</p> <p>Compton scattering hard X-ray source for 10-80 <span class="hlt">keV</span> are under construction using the X-<span class="hlt">band</span> (11.424 GHz) electron linear accelerator and YAG laser at Nuclear Engineering Research laboratory, University of Tokyo. This work is a part of the national project on the development of advanced compact medical accelerators in Japan. National Institute for Radiological Science is the host institute and U.Tokyo and KEK are working for the X-ray source. Main advantage is to produce tunable monochromatic hard (10-80 <span class="hlt">keV</span>) X-rays with the intensities of 108-1010 photons/s (at several stages) and the table-top size. Second important aspect is to reduce noise radiation at a beam dump by adopting the deceleration of electrons after the Compton scattering. This realizes one beamline of a 3rd generation SR source at small facilities without heavy shielding. The final goal is that the linac and laser are installed on the moving gantry. We have designed the X-<span class="hlt">band</span> (11.424 GHz) traveling-wave-type linac for the purpose. Numerical consideration by CAIN code and luminosity calculation are performed to estimate the X-ray yield. X-<span class="hlt">band</span> thermionic-cathode RF-gun and RDS(Round Detuned Structure)-type X-<span class="hlt">band</span> accelerating structure are applied to generate 50 MeV electron beam with 20 pC microbunches (104) for 1 microsecond RF macro-pulse. The X-ray yield by the electron beam and Q-switch Nd:YAG laser of 2 J/10 ns is 107 photons/RF-pulse (108 photons/sec at 10 pps). We design to adopt a technique of laser circulation to increase the X-ray yield up to 109 photons/pulse (1010 photons/s). 50 MW X-<span class="hlt">band</span> klystron and compact modulator have been constructed and now under tuning. The construction of the whole system has started. X-ray generation and medical application will be performed in the early next year.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014SPIE.9144E..1IB','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014SPIE.9144E..1IB"><span>A high-<span class="hlt">energy</span> Compton polarimeter for the POET SMEX mission</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bloser, Peter F.; McConnell, Mark L.; Legere, Jason S.; Ertley, Camden D.; Hill, Joanne E.; Kippen, Marc; Ryan, James M.</p> <p>2014-07-01</p> <p>The primary science goal of the Polarimeters for Energetic Transients (POET) mission is to measure the polarization of gamma-ray bursts over a wide <span class="hlt">energy</span> range, from X rays to soft gamma rays. The higher-<span class="hlt">energy</span> portion of this <span class="hlt">band</span> (50 - 500 <span class="hlt">keV</span>) will be covered by the High <span class="hlt">Energy</span> Polarimeter (HEP) instrument, a non-imaging, wide field of view Compton polarimeter. Incident high-<span class="hlt">energy</span> photons will Compton scatter in low-Z, plastic scintillator detector elements and be subsequently absorbed in high-Z, CsI(Tl) scintillator elements; polarization is detected by measuring an asymmetry in the azimuthal scatter angle distribution. The HEP design is based on our considerable experience with the development and flight of the Gamma-Ray Polarimeter Experiment (GRAPE) balloon payload. We present the design of the POET HEP instrument, which incorporates lessons learned from the GRAPE balloon design and previous work on Explorer proposal efforts, and its expected performance on a two-year SMEX mission.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1356670-grb110721a-extreme-peak-energy-signatures-photosphere','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1356670-grb110721a-extreme-peak-energy-signatures-photosphere"><span>GRB110721A: An extreme peak <span class="hlt">energy</span> and signatures of the photosphere</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Axelsson, M.; Baldini, L.; Barbiellini, G.; ...</p> <p>2012-09-17</p> <p>GRB110721A was observed by the Fermi Gamma-ray Space Telescope using its two instruments, the Large Area Telescope (LAT) and the Gamma-ray Burst Monitor (GBM). This burst consisted of one major emission episode which lasted for ~24.5 s (in the GBM) and had a peak flux of (5.7 ± 0.2) × 10 –5 erg s –1 cm –2. The time-resolved emission spectrum is best modeled with a combination of a <span class="hlt">Band</span> function and a blackbody spectrum. The peak <span class="hlt">energy</span> of the <span class="hlt">Band</span> component was initially 15 ± 2 MeV, which is the highest value ever detected in a GRB. We mademore » this measurement by combining GBM/BGO data with LAT Low <span class="hlt">Energy</span> events to achieve continuous 10-100 MeV coverage. The peak <span class="hlt">energy</span> later decreased as a power law in time with an index of –1.89 ± 0.10. The temperature of the blackbody component also decreased, starting from ~80 <span class="hlt">keV</span>, and the decay showed a significant break after ~2 s. The spectrum provides strong constraints on the standard synchrotron model, indicating that alternative mechanisms may give rise to the emission at these <span class="hlt">energies</span>.« less</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://www.ncbi.nlm.nih.gov/pubmed/25113537','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25113537"><span>A semi-empirical approach to analyze the activities of cylindrical radioactive samples using gamma <span class="hlt">energies</span> from 185 to 1764 <span class="hlt">keV</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Huy, Ngo Quang; Binh, Do Quang</p> <p>2014-12-01</p> <p>This work suggests a method for determining the activities of cylindrical radioactive samples. The self-attenuation factor was applied for providing the self-absorption correction of gamma rays in the sample material. The experimental measurement of a (238)U reference sample and the calculation using the MCNP5 code allow obtaining the semi-empirical formulae of detecting efficiencies for the gamma <span class="hlt">energies</span> ranged from 185 to 1764<span class="hlt">keV</span>. These formulae were used to determine the activities of the (238)U, (226)Ra, (232)Th, (137)Cs and (40)K nuclides in the IAEA RGU-1, IAEA-434, IAEA RGTh-1, IAEA-152 and IAEA RGK-1 radioactive standards. The coincidence summing corrections for gamma rays in the (238)U and (232)Th series were applied. The activities obtained in this work were in good agreement with the reference values. Copyright © 2014 Elsevier Ltd. All rights reserved.</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('https://www.ncbi.nlm.nih.gov/pubmed/29547057','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29547057"><span><span class="hlt">Energy</span>-Specific Optimization of Attenuation Thresholds for Low-<span class="hlt">Energy</span> Virtual Monoenergetic Images in Renal Lesion Evaluation.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Patel, Bhavik N; Farjat, Alfredo; Schabel, Christoph; Duvnjak, Petar; Mileto, Achille; Ramirez-Giraldo, Juan Carlos; Marin, Daniele</p> <p>2018-05-01</p> <p>The purpose of this study was to determine in vitro and in vivo the optimal threshold for renal lesion vascularity at low-<span class="hlt">energy</span> (40-60 <span class="hlt">keV</span>) virtual monoenergetic imaging. A rod simulating unenhanced renal parenchymal attenuation (35 HU) was fitted with a syringe containing water. Three iodinated solutions (0.38, 0.57, and 0.76 mg I/mL) were inserted into another rod that simulated enhanced renal parenchyma (180 HU). Rods were inserted into cylindric phantoms of three different body sizes and scanned with single- and dual-<span class="hlt">energy</span> MDCT. In addition, 102 patients (32 men, 70 women; mean age, 66.8 ± 12.9 [SD] years) with 112 renal lesions (67 nonvascular, 45 vascular) measuring 1.1-8.9 cm underwent single-<span class="hlt">energy</span> unenhanced and contrast-enhanced dual-<span class="hlt">energy</span> CT. Optimal threshold attenuation values that differentiated vascular from nonvascular lesions at 40-60 <span class="hlt">keV</span> were determined. Mean optimal threshold values were 30.2 ± 3.6 (standard error), 20.9 ± 1.3, and 16.1 ± 1.0 HU in the phantom, and 35.9 ± 3.6, 25.4 ± 1.8, and 17.8 ± 1.8 HU in the patients at 40, 50, and 60 <span class="hlt">keV</span>. Sensitivity and specificity for the thresholds did not change significantly between low-<span class="hlt">energy</span> and 70-<span class="hlt">keV</span> virtual monoenergetic imaging (sensitivity, 87-98%; specificity, 90-91%). The AUC from 40 to 70 <span class="hlt">keV</span> was 0.96 (95% CI, 0.93-0.99) to 0.98 (95% CI, 0.95-1.00). Low-<span class="hlt">energy</span> virtual monoenergetic imaging at <span class="hlt">energy</span>-specific optimized attenuation thresholds can be used for reliable characterization of renal lesions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26329204','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26329204"><span>A laboratory 8 <span class="hlt">keV</span> transmission full-field x-ray microscope with a polycapillary as condenser for bright and dark field imaging.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Baumbach, S; Kanngießer, B; Malzer, W; Stiel, H; Wilhein, T</p> <p>2015-08-01</p> <p>This article introduces a laboratory setup of a transmission full-field x-ray microscope at 8 <span class="hlt">keV</span> photon <span class="hlt">energy</span>. The microscope operates in bright and dark field imaging mode with a maximum field of view of 50 μm. Since the illumination geometry determines whether the sample is illuminated homogeneously and moreover, if different imaging methods can be applied, the condenser optic is one of the most significant parts. With a new type of x-ray condenser, a polycapillary optic, we realized bright field imaging and for the first time dark field imaging at 8 <span class="hlt">keV</span> photon <span class="hlt">energy</span> in a laboratory setup. A detector limited spatial resolution of 210 nm is measured on x-ray images of Siemens star test patterns.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018NIMPB.414..107E','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018NIMPB.414..107E"><span>Analytical dependence of effective atomic number on the elemental composition of matter and radiation <span class="hlt">energy</span> in the range 10-1000 <span class="hlt">keV</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Eritenko, A. N.; Tsvetiansky, A. L.; Polev, A. A.</p> <p>2018-01-01</p> <p>In the present paper, a universal analytical dependence of effective atomic number on the composition of matter and radiation <span class="hlt">energy</span> is proposed. This enables one to consider the case of a strong difference in the elemental composition with respect to their atomic numbers over a wide <span class="hlt">energy</span> range. The contribution of photoelectric absorption and incoherent and coherent scattering during the interaction between radiation and matter is considered. For <span class="hlt">energy</span> values over 40 <span class="hlt">keV</span>, the contribution of coherent scattering does not exceed approximately 10% that can be neglected at a further consideration. The effective atomic numbers calculated on the basis of the proposed relationships are compared to the results of calculations based on other methods considered by different authors on the basis of experimental and tabulated data on mass and atomic attenuation coefficients. The examination is carried out for both single-element (e.g., 6C, 14Si, 28Cu, 56Ba, and 82Pb) and multi-element materials. Calculations are performed for W1-xCux alloys (x = 0.35; x = 0.4), PbO, ther moluminescent dosimetry compounds (56Ba, 48Cd, 41Sr, 20Ca, 12Mg, and 11Na), and SO4 in a wide <span class="hlt">energy</span> range. A case with radiation <span class="hlt">energy</span> between the K- and L1-absorption edges is considered for 82Pb, 74W, 56Ba, 48Cd, and 38Sr. This enables to substantially simplify the calculation of the atomic number and will be useful in technical and scientific fields related to the interaction between X-ray/gamma radiation and matter.</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('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('http://adsabs.harvard.edu/abs/2015AIPC.1659d0007M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AIPC.1659d0007M"><span>Mass attenuation coefficient of binderless, pre-treated and tannin-based Rhizophora spp. particleboards using 16.59 - 25.26 <span class="hlt">keV</span> photon <span class="hlt">energy</span> range</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mohd Yusof, Mohd Fahmi; Hamid, Puteri Nor Khatijah Abdul; Bauk, Sabar; Hashim, Rokiah; Tajuddin, Abdul Aziz</p> <p>2015-04-01</p> <p>The Rhizophora spp. particleboards were fabricated using ≤ 104 µm particle size at three different fabrication methods; binderless, steam pre-treated and tannin-added. The mass attenuation coefficient of Rhizophora spp. particleboards were measured using x-ray fluorescent (XRF) photon from niobium, molybdenum, palladium, silver and tin metal plates that provided photon <span class="hlt">energy</span> between 16.59 to 25.26 <span class="hlt">keV</span>. The results were compared to theoretical values for water calculated using photon cross-section database (XCOM).The results showed that all Rhizophora spp. particleboards having mass attenuation coefficient close to calculated XCOM for water. Tannin-added Rizophora spp. particleboard was nearest to calculated XCOM for water with χ2 value of 13.008 followed by binderless Rizophora spp. (25.859) and pre-treated Rizophora spp. (91.941).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018MNRAS.476.1258T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018MNRAS.476.1258T"><span>On the nature of the high-<span class="hlt">energy</span> rollover in 1H 0419-577</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Turner, T. J.; Reeves, J. N.; Braito, V.; Costa, M.</p> <p>2018-05-01</p> <p>A NuSTAR/Swift observation of the luminous Seyfert 1 galaxy 1H 0419-577 taken during 2015 reveals one of the most extreme high-<span class="hlt">energy</span> cut-offs observed to date from an AGN - an origin due to thermal Comptonization would imply a remarkably low coronal temperature kT ˜ 15 <span class="hlt">keV</span>. The low-<span class="hlt">energy</span> peak of the spectrum in the hard X-ray NuSTAR <span class="hlt">band</span>, which peaks before the expected onset of a Compton hump, rules out strong reflection as the origin of the hard excess in this AGN. We show the origin of the high-<span class="hlt">energy</span> rollover is likely due to a combination of both thermal Comptonization and an intrinsically steeper continuum, which is modified by absorption at lower <span class="hlt">energies</span>. Furthermore, modelling the broad-<span class="hlt">band</span> XUV continuum shape as a colour-corrected accretion disc, requires the presence of a variable warm absorber to explain all flux and spectral states of the source, consistent with the previous work on this AGN. While absorber variations produce marked spectral variability in this AGN, consideration of all flux states allows us to isolate a colourless component of variability that may arise from changes in the inner accretion flow, typically at around 10 rg.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/8677807','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/8677807"><span>Absorbed dose in AgBr in direct film for photon <span class="hlt">energies</span> ( < 150 <span class="hlt">keV</span>): relation to optical density. Theoretical calculation and experimental evaluation.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Helmrot, E; Alm Carlsson, G</p> <p>1996-01-01</p> <p> films were found to be proportional to the mean absorbed dose in the AgBr grains calculated according to GREENING'S theory. They were also found to be proportional to the collision kerma in silver bromide (Kc,AgBr) indicating proportionality between Kc,AgBr and the mean absorbed dose in silver bromide. While GREENING'S theory shows that the quotient of the mean absorbed dose in silver bromide and Kc,AgBr varies with photon <span class="hlt">energy</span>, this is not apparent when averaged over the broad (diagnostic) X-ray <span class="hlt">energy</span> spectra used here. Alternatively, proportionality between Kc,AgBr and the mean absorbed dose in silver bromide can be interpreted as resulting from a combination of the SPIERS-CHARLTON theory, valid at low photon <span class="hlt">energies</span> ( < 30 <span class="hlt">keV</span>) and GREENING'S theory, which is strictly valid at <span class="hlt">energies</span> above 50 <span class="hlt">keV</span>. This study shows that the blackening of non-screen films can be related directly to the <span class="hlt">energy</span> absorbed in the AgBr grains of the emulsion layer and that, for the purpose of modelling the imaging chain in intraoral radiography, film response can be represented by Kc,AgBr (at the position of the film) independent of photon <span class="hlt">energy</span>. The importance of taking the complete X-ray <span class="hlt">energy</span> spectrum into full account in deriving Kc,AgBr is clearly demonstrated, showing that the concept of effective <span class="hlt">energy</span> must be used with care.</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://ntrs.nasa.gov/search.jsp?R=20030002464&hterms=sources+energy&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dsources%2Benergy','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20030002464&hterms=sources+energy&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dsources%2Benergy"><span>The BATSE Earth Occultation Catalog of Low <span class="hlt">Energy</span> Gamma Ray Sources</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Harmon, B. A.; WilsonHodge, C. A.; Fishman, G. J.; Paciesas, W.</p> <p>2002-01-01</p> <p>The Burst and Transient Source Experiment (BATSE), aboard the Compton Gamma Ray Observatory (CGRO), provided a record of the hard X-ray/low <span class="hlt">energy</span> gamma ray sky between April 1991 and June 2000. During that time, a catalog of known sources was derived from existing catalogs such as HEAO A-4, as well as new transient sources discovered with BATSE and other X-ray monitors operating in the CGRO era. The Earth Occultation Technique was used to monitor a combination of these sources, mostly galactic, totaling to about 175 objects. The catalog will present the global properties of these sources and their probability of detection (> 10 mCrab, 20-100 <span class="hlt">keV</span>) with BATSE. Systematic errors due to unknown sources or background components are included. Cursory analyses to search for new transients (35-80 mCrab in the 20-100 <span class="hlt">keV</span> <span class="hlt">band</span>) and super-orbital periods in known binary sources are also presented. Whole mission light curves and associated data production/analysis tools are being delivered to the HEASARC for public use.</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/2018PASJ...70...32N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PASJ...70...32N"><span><span class="hlt">Energy</span>-dependent intensity variation of the persistent X-ray emission of magnetars observed with Suzaku</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Nakagawa, Yujin; Ebisawa, Ken; Enoto, Teruaki</p> <p>2018-03-01</p> <p>The emission mechanism of magnetars is still controversial even though various observational and theoretical studies have been made. In order to investigate mechanisms of both the persistent X-ray emission and the burst emission of the magnetars, we propose a model in which the persistent X-ray emission consists of numerous micro-bursts of various sizes. If this model is correct, root mean square (rms) intensity variations of the persistent emission would exceed the values expected from the Poisson distribution. Using Suzaku archive data of 11 magnetars (22 observations), the rms intensity variations were calculated from 0.2 <span class="hlt">keV</span> to 70 <span class="hlt">keV</span>. As a result, we found significant excess rms intensity variations from all 11 magnetars. We suppose that numerous micro-bursts constituting the persistent X-ray emission cause the observed variations, suggesting that the persistent X-ray emission and the burst emission have identical emission mechanisms. In addition, we found that the rms intensity variations clearly increase toward higher <span class="hlt">energy</span> <span class="hlt">bands</span> for four magnetars (six observations). The <span class="hlt">energy</span>-dependent rms intensity variations imply that the soft thermal component and the hard X-ray component are emitted from different regions far apart from each other.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018RScI...89e3301C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018RScI...89e3301C"><span>CVD diamond detector with interdigitated electrode pattern for time-of-flight <span class="hlt">energy</span>-loss measurements of low-<span class="hlt">energy</span> ion bunches</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Cayzac, W.; Pomorski, M.; Blažević, A.; Canaud, B.; Deslandes, D.; Fariaut, J.; Gontier, D.; Lescoute, E.; Marmouget, J. G.; Occelli, F.; Oudot, G.; Reverdin, C.; Sauvestre, J. E.; Sollier, A.; Soullié, G.; Varignon, C.; Villette, B.</p> <p>2018-05-01</p> <p>Ion stopping experiments in plasma for beam <span class="hlt">energies</span> of few hundred <span class="hlt">keV</span> per nucleon are of great interest to benchmark the stopping-power models in the context of inertial confinement fusion and high-<span class="hlt">energy</span>-density physics research. For this purpose, a specific ion detector on chemical-vapor-deposition diamond basis has been developed for precise time-of-flight measurements of the ion <span class="hlt">energy</span> loss. The electrode structure is interdigitated for maximizing its sensitivity to low-<span class="hlt">energy</span> ions, and it has a finger width of 100 μm and a spacing of 500 μm. A short single α-particle response is obtained, with signals as narrow as 700 ps at full width at half maximum. The detector has been tested with α-particle bunches at a 500 <span class="hlt">keV</span> per nucleon <span class="hlt">energy</span>, showing an excellent time-of-flight resolution down to 20 ps. In this way, beam <span class="hlt">energy</span> resolutions from 0.4 <span class="hlt">keV</span> to a few <span class="hlt">keV</span> have been obtained in an experimental configuration using a 100 μg/cm2 thick carbon foil as an <span class="hlt">energy</span>-loss target and a 2 m time-of-flight distance. This allows a highly precise beam <span class="hlt">energy</span> measurement of δE/E ≈ 0.04%-0.2% and a resolution on the <span class="hlt">energy</span> loss of 0.6%-2.5% for a fine testing of stopping-power models.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1996APS..DPP..5E01G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1996APS..DPP..5E01G"><span>Efficient Production of 4-<span class="hlt">KeV</span> X Rays from Laser-Heated Xe Gas = Confined Within a Hohlraum</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Grun, Jacob; Suter, Larry J.; Back, Christina A.; Decker, Chris; Kauffman, Robert L.; Davis, John F.</p> <p>1996-11-01</p> <p>Clean (debris-free) and efficient multi-kilovolt x-ray sources are needed for irradiating large military test objects and for use as backlighters in future Inertial Confinement Fusion experiments. Laser-plasma x-ray sources are particularly attractive for these uses since their spectrum can be controlled by proper choice of plasma material and laser intensity; and because many laser-plasma sources can be designed to produce little or no particulate debris. We report on an experiment in which we measured the production-efficiency, spectrum, and time history of 1-4 <span class="hlt">KeV</span> x-rays from beryllium hohlraums which were filled with 1 and 2 atm of Xe gas and then irradiated by a 2-nsec pulse from the NOVA laser. It is predicted that 17be converted into > 4<span class="hlt">KeV</span> x rays and 30history of >4<span class="hlt">KeV</span> part of the spectrum is predicted to exhibit a dip in intensity whose depth and location vary with fill pressure and hohlraum size.. We also measured the debris produced by these sources. Work supported by the Defense Special Weapons Agency and the U.S. Department of <span class="hlt">Energy</span> at LLNL under W-7405-ENG-48.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFMSM22C..06Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFMSM22C..06Y"><span>The characteristic pitch angle distributions of 1 eV to 600 <span class="hlt">keV</span> protons near the equator based on Van Allen Probes observations</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yue, C.; Bortnik, J.; Thorne, R. M.; Ma, Q.; An, X.; Chappell, C. R.; Gerrard, A. J.; Lanzerotti, L. J.; Shi, Q.</p> <p>2017-12-01</p> <p>Understanding the source and loss processes of various plasma populations is greatly aided by having accurate knowledge of their pitch angle distributions (PADs). Here, we statistically analyze 1 eV to 600 <span class="hlt">keV</span> hydrogen (H+) PADs near the geomagnetic equator in the inner magnetosphere based on Van Allen Probes measurements, to comprehensively investigate how the H+ PADs vary with different <span class="hlt">energies</span>, magnetic local times (MLTs), L-shells, and geomagnetic conditions. Our survey clearly indicates four distinct populations with different PADs: (1) a pancake distribution of the plasmaspheric H+ at low L-shells except for dawn sector; (2) a bi-directional field-aligned distribution of the warm plasma cloak; (3) pancake or isotropic distributions of ring current H+; (4) radiation belt particles show pancake, butterfly and isotropic distributions depending on their <span class="hlt">energy</span>, MLT and L-shell. Meanwhile, the pancake distribution of ring current H+ moves to lower <span class="hlt">energies</span> as L-shell increases which is primarily caused by adiabatic transport. Furthermore, energetic H+ (> 10 <span class="hlt">keV</span>) PADs become more isotropic following the substorm injections, indicating wave-particle interactions. The radiation belt H+ butterfly distributions are identified in a narrow <span class="hlt">energy</span> range of 100 < E < 400 <span class="hlt">keV</span> at large L (L > 5), which are less significant during quiet times and extend from dusk to dawn sector through midnight during substorms. The different PADs near the equator provide clues of the underlying physical processes that produce the dynamics of these different populations.</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> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2005NIMPA.545..744O','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2005NIMPA.545..744O"><span>Avalanche photodiode for measurement of low-<span class="hlt">energy</span> electrons</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ogasawara, K.; Asamura, K.; Mukai, T.; Saito, Y.</p> <p>2005-06-01</p> <p>We report on the performance of an Avalanche Photodiode (APD) produced by Hamamatsu Photonics Co. Ltd. (Type Z7966-20) for measurements of low <span class="hlt">energy</span> electrons. We have set up an electron gun, which can generate a 1-20 <span class="hlt">keV</span> electron beam impinging onto the APD in a vacuum chamber. The result shows that the pulse height distribution (PHD) of the APD signal exhibits a significant peak for electrons with <span class="hlt">energies</span> above 8 <span class="hlt">keV</span>, and the variation of the PHD peak shows a good linearity with the <span class="hlt">energy</span> of incident electrons. The <span class="hlt">energy</span> resolution is quite good, though it slightly depends on the electron <span class="hlt">energy</span>. In the case of low-<span class="hlt">energies</span> (lower than 10 <span class="hlt">keV</span>), the pulse height distribution has a characteristic tail on the low <span class="hlt">energy</span> side, and the <span class="hlt">energy</span> resolution becomes a little worse. The position of the peak appears on a slightly lower channel than is expected from data at higher <span class="hlt">energies</span> (near 20 <span class="hlt">keV</span>). Qualitatively, the low-<span class="hlt">energy</span> tail is caused by the dead-layer on the surface of the device. The nonlinearity and the worse resolution of the peaks for higher <span class="hlt">energy</span> electrons may have resulted from a space-charge effect due to created e-h pairs. For a quantitative understanding, we have made a Monte Carlo particle simulation of charge transport and collection inside the APD.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22482746-laboratory-kev-transmission-full-field-ray-microscope-polycapillary-condenser-bright-dark-field-imaging','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22482746-laboratory-kev-transmission-full-field-ray-microscope-polycapillary-condenser-bright-dark-field-imaging"><span>A laboratory 8 <span class="hlt">keV</span> transmission full-field x-ray microscope with a polycapillary as condenser for bright and dark field imaging</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>Baumbach, S., E-mail: baumbach@rheinahrcampus.de; Wilhein, T.; Kanngießer, B.</p> <p>2015-08-15</p> <p>This article introduces a laboratory setup of a transmission full-field x-ray microscope at 8 <span class="hlt">keV</span> photon <span class="hlt">energy</span>. The microscope operates in bright and dark field imaging mode with a maximum field of view of 50 μm. Since the illumination geometry determines whether the sample is illuminated homogeneously and moreover, if different imaging methods can be applied, the condenser optic is one of the most significant parts. With a new type of x-ray condenser, a polycapillary optic, we realized bright field imaging and for the first time dark field imaging at 8 <span class="hlt">keV</span> photon <span class="hlt">energy</span> in a laboratory setup. A detectormore » limited spatial resolution of 210 nm is measured on x-ray images of Siemens star test patterns.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20170003457&hterms=Andromeda&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3DAndromeda','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20170003457&hterms=Andromeda&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3DAndromeda"><span>Searching for the 3.5 <span class="hlt">keV</span> Line in the Stacked Suzaku Observations of Galaxy Clusters</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Bulbul, Esra; Markevitch, Maxim; Foster, Adam; Miller, Eric; Bautz, Mark; Lowenstein, Mike; Randall, Scott W.; Smith, Randall K.</p> <p>2016-01-01</p> <p>We perform a detailed study of the stacked Suzaku observations of 47 galaxy clusters, spanning a redshift range of 0.01-0.45, to search for the unidentified 3.5 <span class="hlt">keV</span> line. This sample provides an independent test for the previously detected line. We detect a 2sigma-significant spectral feature at 3.5 <span class="hlt">keV</span> in the spectrum of the full sample. When the sample is divided into two subsamples (cool-core and non-cool core clusters), the cool-core subsample shows no statistically significant positive residuals at the line <span class="hlt">energy</span>. A very weak (approx. 2sigma confidence) spectral feature at 3.5 <span class="hlt">keV</span> is permitted by the data from the non-cool-core clusters sample. The upper limit on a neutrino decay mixing angle of sin(sup 2)(2theta) = 6.1 x 10(exp -11) from the full Suzaku sample is consistent with the previous detections in the stacked XMM-Newton sample of galaxy clusters (which had a higher statistical sensitivity to faint lines), M31, and Galactic center, at a 90% confidence level. However, the constraint from the present sample, which does not include the Perseus cluster, is in tension with previously reported line flux observed in the core of the Perseus cluster with XMM-Newton and Suzaku.</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 addition, 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.osti.gov/biblio/22075717-optima-mdxt-high-throughput-kev-mid-dose-implanter','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22075717-optima-mdxt-high-throughput-kev-mid-dose-implanter"><span>Optima MDxt: A high throughput 335 <span class="hlt">keV</span> mid-dose implanter</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>Eisner, Edward; David, Jonathan; Justesen, Perry</p> <p>2012-11-06</p> <p>The continuing demand for both <span class="hlt">energy</span> purity and implant angle control along with high wafer throughput drove the development of the Axcelis Optima MDxt mid-dose ion implanter. The system utilizes electrostatic scanning, an electrostatic parallelizing lens and an electrostatic <span class="hlt">energy</span> filter to produce energetically pure beams with high angular integrity. Based on field proven components, the Optima MDxt beamline architecture offers the high beam currents possible with singly charged species including arsenic at <span class="hlt">energies</span> up to 335 <span class="hlt">keV</span> as well as large currents from multiply charged species at <span class="hlt">energies</span> extending over 1 MeV. Conversely, the excellent <span class="hlt">energy</span> filtering capability allowsmore » high currents at low beam <span class="hlt">energies</span>, since it is safe to utilize large deceleration ratios. This beamline is coupled with the >500 WPH capable endstation technology used on the Axcelis Optima XEx high <span class="hlt">energy</span> ion implanter. The endstation includes in-situ angle measurements of the beam in order to maintain excellent beam-to-wafer implant angle control in both the horizontal and vertical directions. The Optima platform control system provides new generation dose control system that assures excellent dosimetry and charge control. This paper will describe the features and technologies that allow the Optima MDxt to provide superior process performance at the highest wafer throughput, and will provide examples of the process performance achievable.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017PhRvC..95e4308W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017PhRvC..95e4308W"><span>Three-quasiparticle isomer in 173Ta and the excitation <span class="hlt">energy</span> dependence of K -forbidden transition rates</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wood, R. T.; Walker, P. M.; Lane, G. J.; Carroll, R. J.; Cullen, D. M.; Dracoulis, G. D.; Hota, S. S.; Kibédi, T.; Palalani, N.; Podolyák, Zs.; Reed, M. W.; Schiffl, K.; Wright, A. M.</p> <p>2017-05-01</p> <p>Using the 168Er(10B,5 n ) reaction at a beam <span class="hlt">energy</span> of 68 MeV, new data have been obtained for the population and decay of a T1 /2=148 ns, Kπ=21 /2- three-quasiparticle isomer at 1717 <span class="hlt">keV</span> in 173Ta. Revised decay <span class="hlt">energies</span> and intensities have been determined, together with newly observed members of a rotational <span class="hlt">band</span> associated with the isomer. By comparison with other isomers in the A ≈180 deformed region, the 173Ta isomer properties help to specify the key degrees of freedom that determine K -forbidden transition rates. In particular, when all three quasiparticles are of the same nucleon type, there is a strong dependence of the E 2 reduced hindrance factor on the isomer excitation <span class="hlt">energy</span>.</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 addition, 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('http://adsabs.harvard.edu/abs/2005MNRAS.362.1371L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2005MNRAS.362.1371L"><span>XMM-Newton 13H deep field - I. X-ray sources</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Loaring, N. S.; Dwelly, T.; Page, M. J.; Mason, K.; McHardy, I.; Gunn, K.; Moss, D.; Seymour, N.; Newsam, A. M.; Takata, T.; Sekguchi, K.; Sasseen, T.; Cordova, F.</p> <p>2005-10-01</p> <p>We present the results of a deep X-ray survey conducted with XMM-Newton, centred on the UK ROSAT13H deep field area. This region covers 0.18 deg2, and is the first of the two areas covered with XMM-Newton as part of an extensive multiwavelength survey designed to study the nature and evolution of the faint X-ray source population. We have produced detailed Monte Carlo simulations to obtain a quantitative characterization of the source detection procedure and to assess the reliability of the resultant sourcelist. We use the simulations to establish a likelihood threshold, above which we expect less than seven (3 per cent) of our sources to be spurious. We present the final catalogue of 225 sources. Within the central 9 arcmin, 68 per cent of source positions are accurate to 2 arcsec, making optical follow-up relatively straightforward. We construct the N(>S) relation in four <span class="hlt">energy</span> <span class="hlt">bands</span>: 0.2-0.5, 0.5-2, 2-5 and 5-10 <span class="hlt">keV</span>. In all but our highest <span class="hlt">energy</span> <span class="hlt">band</span> we find that the source counts can be represented by a double power law with a bright-end slope consistent with the Euclidean case and a break around 10-14yergcm-2s-1. Below this flux, the counts exhibit a flattening. Our source counts reach densities of 700, 1300, 900 and 300 deg-2 at fluxes of 4.1 × 10-16,4.5 × 10-16,1.1 × 10-15 and 5.3 × 10-15ergcm-2s-1 in the 0.2-0.5, 0.5-2, 2-5 and 5-10 <span class="hlt">keV</span> <span class="hlt">energy</span> <span class="hlt">bands</span>, respectively. We have compared our source counts with those in the two Chandra deep fields and Lockman hole, and found our source counts to be amongst the highest of these fields in all <span class="hlt">energy</span> <span class="hlt">bands</span>. We resolve >51 per cent (>50 per cent) of the X-ray background emission in the 1-2 <span class="hlt">keV</span> (2-5 <span class="hlt">keV</span>) <span class="hlt">energy</span> <span class="hlt">bands</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015RaPC..109...89L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015RaPC..109...89L"><span>Studies on mass <span class="hlt">energy</span>-absorption coefficients and effective atomic <span class="hlt">energy</span>-absorption cross sections for carbohydrates</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ladhaf, Bibifatima M.; Pawar, Pravina P.</p> <p>2015-04-01</p> <p>We measured here the mass attenuation coefficients (μ/ρ) of carbohydrates, Esculine (C15H16O9), Sucrose (C12H22O11), Sorbitol (C6H14O6), D-Galactose (C6H12O6), Inositol (C6H12O6), D-Xylose (C5H10O5) covering the <span class="hlt">energy</span> range from 122 <span class="hlt">keV</span> up to 1330 <span class="hlt">keV</span> photon <span class="hlt">energies</span> by using gamma ray transmission method in a narrow beam good geometry set-up. The gamma-rays were detected using NaI(Tl) scintillation detection system with a resolution of 8.2% at 662 <span class="hlt">keV</span>. The attenuation coefficient data were then used to obtain the total attenuation cross-section (σtot), molar extinction coefficients (ε), mass-<span class="hlt">energy</span> absorption coefficients (μen/ρ) and effective (average) atomic <span class="hlt">energy</span>-absorption cross section (σa,en) of the compounds. These values are found to be in good agreement with the theoretical values calculated based on XCOM data.</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/2017PhRvC..95b4329A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017PhRvC..95b4329A"><span>Lifetime measurements in 162Dy</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Aprahamian, A.; Lesher, S. R.; Casarella, C.; Börner, H. G.; Jentschel, M.</p> <p>2017-02-01</p> <p>Background: The nature of oscillations or excitations around the equilibrium deformed nuclear shape remains an open question in nuclear structure. The 162Dy nucleus is one of the most extensively studied nuclei with the (n ,γ ), (n ,e- ), (α ,2 n ) reactions and most recently the (p ,t ) pickup reaction adding 11 0+ states to an excitation <span class="hlt">energy</span> of 2.8 MeV to an already-well-developed level scheme. However, a major shortfall for a better understanding of the nature of the plethora of <span class="hlt">bands</span> and levels in this nucleus has been the lack of lifetime measurements. Purpose: To determine the character of the low-lying excited <span class="hlt">bands</span> in this 162Dy nucleus, we set out to measure the level lifetimes. Method: Lifetimes were measured in the 162Dy nucleus following neutron capture using the Gamma-Ray-Induced Doppler (GRID) broadening technique at the Institut Laue-Langevin in Grenoble, France. Results: In total, we have measured the lifetimes of 12 levels belonging to a number of excited positive- and negative-parity <span class="hlt">bands</span> in the low-lying spectrum of the 162Dy nucleus. The lifetime of the Kπ=2+ bandhead at 888.16 <span class="hlt">keV</span> was previously measured. We confirm this value and measure lifetimes of the 3+ and 4+ members of this <span class="hlt">band</span> yielding B (E 2 ) values that are consistent with a single γ -vibrational phonon of several Weisskopf units. The first excited Kπ=4+ <span class="hlt">band</span>, with a bandhead at 1535.66 <span class="hlt">keV</span>, is strongly connected to the Kπ=2+ <span class="hlt">band</span> with enhanced collective B (E 2 ) values and it is consistent with a double phonon vibrational (γ γ ) excitation. Lifetime of Kπ=0+ <span class="hlt">band</span> members have also been measured, including the 4Kπ=02+ state at 1574.29 <span class="hlt">keV</span> and the 2Kπ= 03+ state at 1728.31 <span class="hlt">keV</span>. This latter state also displays the characteristics of a double phonon excitation built on the Kπ=2+ <span class="hlt">band</span>. Conclusions: We discuss our findings in terms of the presence or absence of collective quadrupole and octupole vibrational excitations. We find two positive-parity excited <span class="hlt">bands</span> at 1535</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20080031148&hterms=Skinner&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3DSkinner','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20080031148&hterms=Skinner&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3DSkinner"><span>Asymmetric 511 <span class="hlt">keV</span> Positron Annihilation Line Emission from the Inner Galactic Disk</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Skinner, Gerry; Weidenspointner, Georg; Jean, Pierre; Knodlseder, Jurgen; Ballmoos, Perer von; Bignami, Giovanni; Diehl, Roland; Strong, Andrew; Cordier, Bertrand; Schanne, Stephane; <a style="text-decoration: none; " href="javascript:void(0); " onClick="displayelement('author_20080031148'); toggleEditAbsImage('author_20080031148_show'); toggleEditAbsImage('author_20080031148_hide'); "> <img style="display:inline; width:12px; height:12px; " src="images/arrow-up.gif" width="12" height="12" border="0" alt="hide" id="author_20080031148_show"> <img style="width:12px; height:12px; display:none; " src="images/arrow-down.gif" width="12" height="12" border="0" alt="hide" id="author_20080031148_hide"></p> <p>2008-01-01</p> <p>A recently reported asymmetry in the 511 <span class="hlt">keV</span> gamma-ray line emission from the inner galactic disk is unexpected and mimics an equally unexpected one in the distribution of LMXBs seen at hard X-ray <span class="hlt">energies</span>. A possible conclusion is that LMXBs are an important source of the positrons whose annihilation gives rise to the line. We will discuss these results, their statistical significance and that of any link between the two. The implication of any association between LMXBs and positrons for the strong annihilation radiation from the galactic bulge will be reviewed.</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('http://adsabs.harvard.edu/abs/2016AGUFM.P13A1911G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFM.P13A1911G"><span>The Martian diffuse aurora: Monte Carlo simulations and comparison with IUVS-MAVEN observations</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gerard, J. C. M. C.; Soret, L.; Schneider, N. M.; Shematovich, V.; Bisikalo, D.; Bougher, S. W.; Jain, S.; Lillis, R. J.; Mitchell, D. L.; Jakosky, B. M.; Deighan, J.; Larson, D. E.</p> <p>2016-12-01</p> <p>A new type of Martian aurora, characterized by an extended spatial distribution, an altitude lower than the discrete aurora and electron precipitation up to 200 <span class="hlt">keV</span> has been observed following solar activity on several occasions with the IUVS on board the MAVEN spacecraft. We describe the results of Monte Carlo simulations of the production of several ultraviolet and visible auroral emissions for initial electron <span class="hlt">energies</span> from 0.1 to 200 <span class="hlt">keV</span>. These include the CO2+ ultraviolet doublet (UVD) at 288.3 and 289.6 nm and the Fox-Duffendack-Barker (FDB) <span class="hlt">bands</span>, CO Cameron and Fourth Positive <span class="hlt">bands</span>, OI 130.4 and 297.2 nm and CI 156.1 nm and 165.7 nm multiplets. We calculate the nadir and limb intensities of several of these emissions for a unit precipitated <span class="hlt">energy</span> flux. Our results indicate that electrons in the range 100-200 <span class="hlt">keV</span> produce maximum CO2+ UVD emission near 75 km. We combine SWEA and SEP electron <span class="hlt">energy</span> spectra measured during diffuse aurora to calculate the volume emission rates and compare with IUVS observations of the emission limb profiles. The strongest predicted emissions are the CO2+ FDB, UVD and the CO Cameron <span class="hlt">bands</span>. The metastable a 3Π state which radiates the Cameron <span class="hlt">bands</span> is deactivated by collisions below 110 km. As a consequence, we show that the CO2+ UVD to the Cameron <span class="hlt">bands</span> ratio increases at low altitude in the energetic diffuse aurora.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2009A%26A...494...49P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009A%26A...494...49P"><span>High <span class="hlt">energy</span> variability of 3C 273 during the AGILE multiwavelength campaign of December 2007-January 2008</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Pacciani, L.; Donnarumma, I.; Vittorini, V.; D'Ammando, F.; Fiocchi, M. T.; Impiombato, D.; Stratta, G.; Verrecchia, F.; Bulgarelli, A.; Chen, A. W.; Giuliani, A.; Longo, F.; Pucella, G.; Vercellone, S.; Tavani, M.; Argan, A.; Barbiellini, G.; Boffelli, F.; Caraveo, P. A.; Cattaneo, P. W.; Cocco, V.; Costa, E.; Del Monte, E.; Di Cocco, G.; Evangelista, Y.; Feroci, M.; Froysland, T.; Fuschino, F.; Galli, M.; Gianotti, F.; Labanti, C.; Lapshov, I.; Lazzarotto, F.; Lipari, P.; Marisaldi, M.; Mereghetti, S.; Morselli, A.; Pellizzoni, A.; Perotti, F.; Picozza, P.; Prest, M.; Rapisarda, M.; Soffitta, P.; Trifoglio, M.; Tosti, G.; Trois, A.; Vallazza, E.; Zanello, D.; Antonelli, L. A.; Colafrancesco, S.; Cutini, S.; Gasparrini, D.; Giommi, P.; Pittori, C.; Salotti, L.</p> <p>2009-01-01</p> <p>Context: We report the results of a 3-week multi-wavelength campaign targeting the flat spectrum radio quasar 3C 273 carried out with the AGILE gamma-ray mission, covering the 30 MeV-50 GeV and 18-60 <span class="hlt">keV</span>, the REM observatory (covering the near-IR and optical), Swift (near-UV/Optical, 0.2-10 <span class="hlt">keV</span> and 15-50 <span class="hlt">keV</span>), INTEGRAL (3-200 <span class="hlt">keV</span>) and Rossi XTE (2-12 <span class="hlt">keV</span>). This is the first observational campaign including gamma-ray data, after the last EGRET observations, more than 8 years ago. Aims: This campaign has been organized by the AGILE team with the aim of observing, studying and modelling the broad <span class="hlt">band</span> <span class="hlt">energy</span> spectrum of the source, and its variability on a week timescale, testing the emission models describing the spectral <span class="hlt">energy</span> distribution of this source. Methods: Our study was carried out using simultaneous light curves of the source flux from all the involved instruments, in the different <span class="hlt">energy</span> ranges, to search for correlated variability. Then a time-resolved spectral <span class="hlt">energy</span> distribution was used for a detailed physical modelling of the emission mechanisms. Results: The source was detected in gamma-rays only in the second week of our campaign, with a flux comparable to the level detected by EGRET in June 1991. We found an indication of a possible anti-correlation between the emission at gamma-rays and at soft and hard X-rays, supported by the complete set of instruments. Instead, optical data do not show short term variability, as expected for this source. Only in two preceding EGRET observations (in 1993 and 1997) 3C 273 showed intra-observation variability in gamma-rays. In the 1997 observation, flux variation in gamma-rays was associated with a synchrotron flare. The <span class="hlt">energy</span>-density spectrum with almost simultaneous data partially covers the regions of synchrotron emission, the big blue bump, and the inverse-Compton. We adopted a leptonic model to explain the hard X/gamma-ray emissions, although from our analysis hadronic models cannot be ruled out. In the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015SpWea..13...16G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015SpWea..13...16G"><span>Nowcast model for low-<span class="hlt">energy</span> electrons in the inner magnetosphere</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ganushkina, N. Yu.; Amariutei, O. A.; Welling, D.; Heynderickx, D.</p> <p>2015-01-01</p> <p>We present the nowcast model for low-<span class="hlt">energy</span> (<200 <span class="hlt">keV</span>) electrons in the inner magnetosphere, which is the version of the Inner Magnetosphere Particle Transport and Acceleration Model (IMPTAM) for electrons. Low-<span class="hlt">energy</span> electron fluxes are very important to specify when hazardous satellite surface-charging phenomena are considered. The presented model provides the low-<span class="hlt">energy</span> electron flux at all L shells and at all satellite orbits, when necessary. The model is driven by the real-time solar wind and interplanetary magnetic field (IMF) parameters with 1 h time shift for propagation to the Earth's magnetopause and by the real time Dst index. Real-time geostationary GOES 13 or GOES 15 (whenever each is available) data on electron fluxes in three <span class="hlt">energies</span>, such as 40 <span class="hlt">keV</span>, 75 <span class="hlt">keV</span>, and 150 <span class="hlt">keV</span>, are used for comparison and validation of IMPTAM running online. On average, the model provides quite reasonable agreement with the data; the basic level of the observed fluxes is reproduced. The best agreement between the modeled and the observed fluxes are found for <100 <span class="hlt">keV</span> electrons. At the same time, not all the peaks and dropouts in the observed electron fluxes are reproduced. For 150 <span class="hlt">keV</span> electrons, the modeled fluxes are often smaller than the observed ones by an order of magnitude. The normalized root-mean-square deviation is found to range from 0.015 to 0.0324. Though these metrics are buoyed by large standard deviations, owing to the dynamic nature of the fluxes, they demonstrate that IMPTAM, on average, predicts the observed fluxes satisfactorily. The computed binary event tables for predicting high flux values within each 1 h window reveal reasonable hit rates being 0.660-0.318 for flux thresholds of 5 ·104-2 ·105 cm-2 s-1 sr-1 <span class="hlt">keV</span>-1 for 40 <span class="hlt">keV</span> electrons, 0.739-0.367 for flux thresholds of 3 ·104-1 ·105 cm-2 s-1 sr-1 <span class="hlt">keV</span>-1 for 75 <span class="hlt">keV</span> electrons, and 0.485-0.438 for flux thresholds of 3 ·103-3.5 ·103 cm-2 s-1 sr-1 <span class="hlt">keV</span>-1 for 150 <span class="hlt">keV</span> electrons but rather small Heidke</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://ntrs.nasa.gov/search.jsp?R=20050182675&hterms=Ford&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3DFord','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20050182675&hterms=Ford&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3DFord"><span>Discovery of Oxygen Kalpha X-ray Emission from the Rings of Saturn</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Bhardwaj, Anil; Elsner, Ronald F.; Waite, J. Hunter, Jr.; Gladstone, G Randall; Cravens, Thomas E.; Ford, Peter G.</p> <p>2005-01-01</p> <p>Using the Advanced CCD Imaging Spectrometer (ACIS), the Chandra X-ray Observatory (CXO) observed the Saturnian system for one rotation of the planet (approx.37 ks) on 20 January, 2004, and again on 26-27 January, 2004. In this letter we report the detection of X-ray emission from the rings of Saturn. The X-ray spectrum from the rings is dominated by emission in a narrow (approx.130 eV wide) <span class="hlt">energy</span> <span class="hlt">band</span> centered on the atomic oxygen Ka fluorescence line at 0.53 <span class="hlt">keV</span>. The X-ray power emitted from the rings in the 0.49-0.62 <span class="hlt">keV</span> <span class="hlt">band</span> is about one-third of that emitted from Saturn disk in the photon <span class="hlt">energy</span> range 0.24-2.0 <span class="hlt">keV</span>. Our analysis also finds a clear detection of X-ray emission from the rings in the 0.49-0.62 <span class="hlt">keV</span> <span class="hlt">band</span> in an earlier (14-15 April, 2003) Chandra ACIS observation of Saturn. Fluorescent scattering of solar X-rays from oxygen atoms in the H20 icy ring material is the likely source mechanism for ring X-rays, consistent with the scenario of solar photo-production of a tenuous ring oxygen atmosphere and ionosphere recently discovered by Cassini.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018NIMPA.893...26K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018NIMPA.893...26K"><span><span class="hlt">Energy</span> dependent features of X-ray signals in a GridPix detector</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Krieger, C.; Kaminski, J.; Vafeiadis, T.; Desch, K.</p> <p>2018-06-01</p> <p>We report on the calibration of an argon/isobutane (97.7%/2.3%)-filled GridPix detector with soft X-rays (277 eV to 8 <span class="hlt">keV</span>) using the variable <span class="hlt">energy</span> X-ray source of the CAST Detector Lab at CERN. We study the linearity and <span class="hlt">energy</span> resolution of the detector using both the number of pixels hit and the total measured charge as <span class="hlt">energy</span> measures. For the latter, the <span class="hlt">energy</span> resolution σE / E is better than 10% (20%) for <span class="hlt">energies</span> above 2 <span class="hlt">keV</span> (0.5 <span class="hlt">keV</span>). Several characteristics of the recorded events are studied.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22253291-study-kev-ray-emission-from-high-intensity-femtosecond-laser-produced-plasma','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22253291-study-kev-ray-emission-from-high-intensity-femtosecond-laser-produced-plasma"><span>Study of 1–8 <span class="hlt">keV</span> K-α x-ray emission from high intensity femtosecond laser produced plasma</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>Arora, V., E-mail: arora@rrcat.gov.in; Naik, P. A.; Chakera, J. A.</p> <p>2014-04-15</p> <p>We report an experimental study on the optimization of a laser plasma based x-ray source of ultra-short duration K-α line radiation. The interaction of pulses from a CPA based Ti:sapphire laser (10 TW, 45 fs, 10 Hz) system with magnesium, titanium, iron and copper solid target generates bright 1-8 <span class="hlt">keV</span> K-α x-ray radiation. The x-ray yield was optimized with the laser pulse duration (at fixed fluence) which is varied in the range of 45 fs to 1.4 ps. It showed a maximum at laser pulse duration of ∼740 fs, 420 fs, 350 and 250 fs for Mg (1.3 <span class="hlt">keV</span>), Timore » (4.5 <span class="hlt">keV</span>), Fe (6.4 <span class="hlt">keV</span>) and Cu (8.05 <span class="hlt">keV</span>) respectively. The x-ray yield is observed to be independent of the sign of the chirp. The scaling of the K-α yield (I{sub x} ∝ I{sub L}{sup β}) for 45 fs and optimized pulse duration were measured for laser intensities in the region of 3 × 10{sup 14} – 8 × 10{sup 17}. The x-ray yield shows a much faster scaling exponent β = 1.5, 2.1, 2.4 and 2.6 for Mg, Ti, Fe and Cu respectively at optimized pulse duration compared to scaling exponent of 0.65, 1.3, 1.5, and 1.7 obtained for 45 fs duration laser pulses. The laser to x-ray <span class="hlt">energy</span> conversion efficiencies obtained for different target materials are η{sub Mg} = 1.2 × 10{sup −5}, η{sub Ti} = 3.1 × 10{sup −5}, η{sub Fe} = 2.7 × 10{sup −5}, η{sub Cu} = 1.9 × 10{sup −5}. The results have been explained from the efficient generation of optimal <span class="hlt">energy</span> hot electrons at longer laser pulse duration. The faster scaling observed at optimal pulse duration indicates that the x-ray source is generated at the target surface and saturation of x-ray emission would appear at larger laser fluence. An example of utilization of the source for measurement of shock-wave profiles in a silicon crystal by time resolved x-ray diffraction is also presented.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25430181','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25430181"><span>Compact focusing spectrometer: visible (1 eV) to hard x-rays (200 <span class="hlt">keV</span>).</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Baronova, E O; Stepanenko, A M; Pereira, N R</p> <p>2014-11-01</p> <p>A low-cost spectrometer that covers a wide range of photon <span class="hlt">energies</span> can be useful to teach spectroscopy, and for simple, rapid measurements of the photon spectrum produced by small plasma devices. The spectrometer here achieves its wide range, nominally from 1 eV to 200 <span class="hlt">keV</span>, with a series of spherically and cylindrically bent gratings or crystals that all have the same shape and the same radius of curvature; they are complemented by matching apertures and diagnostics on the Rowland circle that serves as the circular part of the spectrometer's vacuum vessel. Spectral lines are easily identified with software that finds their positions from the dispersion of each diffractive element and the known <span class="hlt">energies</span> of the lines.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018AIPA....8a5121Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AIPA....8a5121Z"><span>Effects of electronic excitation in 150 <span class="hlt">keV</span> Ni ion irradiation of metallic systems</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zarkadoula, Eva; Samolyuk, German; Weber, William J.</p> <p>2018-01-01</p> <p>We use the two-temperature model in molecular dynamic simulations of 150 <span class="hlt">keV</span> Ni ion cascades in nickel and nickel-based alloys to investigate the effect of the <span class="hlt">energy</span> exchange between the atomic and the electronic systems during the primary stages of radiation damage. We find that the electron-phonon interactions result in a smaller amount of defects and affect the cluster formation, resulting in smaller clusters. These results indicate that ignoring the local heating due to the electrons results in the overestimation of the amount of damage and the size of the defect clusters. A comparison of the average defect production to the Norgett-Robinson-Torrens (NRT) prediction over a range of <span class="hlt">energies</span> is provided.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010SPIE.7732E..1LN','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010SPIE.7732E..1LN"><span>The hard x-ray imager onboard IXO</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Nakazawa, Kazuhiro; Takahashi, Tadayuki; Limousin, Olivier; Kokubun, Motohide; Watanabe, Shin; Laurent, Philippe; Arnaud, Monique; Tajima, Hiroyasu</p> <p>2010-07-01</p> <p>The Hard X-ray Imager (HXI) is one of the instruments onboard International X-ray Observatory (IXO), to be launched into orbit in 2020s. It covers the <span class="hlt">energy</span> <span class="hlt">band</span> of 10-40 <span class="hlt">keV</span>, providing imaging-spectroscopy with a field of view of 8 x 8 arcmin2. The HXI is attached beneath the Wide Field Imager (WFI) covering 0.1-15 <span class="hlt">keV</span>. Combined with the super-mirror coating on the mirror assembly, this configuration provides observation of X-ray source in wide <span class="hlt">energy</span> <span class="hlt">band</span> (0.1-40.0 <span class="hlt">keV</span>) simultaneously, which is especially important for varying sources. The HXI sensor part consists of the semiconductor imaging spectrometer, using Si in the medium <span class="hlt">energy</span> detector and CdTe in the high <span class="hlt">energy</span> detector as its material, and an active shield covering its back to reduce background in orbit. The HXI technology is based on those of the Japanese-lead new generation X-ray observatory ASTRO-H, and partly from those developed for Simbol-X. Therefore, the technological development is in good progress. In the IXO mission, HXI will provide a major assets to identify the nature of the object by penetrating into thick absorbing materials and determined the inherent spectral shape in the <span class="hlt">energy</span> <span class="hlt">band</span> well above the structure around Fe-K lines and edges.</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/21530198','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21530198"><span>Low-<span class="hlt">energy</span> photons in high-<span class="hlt">energy</span> photon fields--Monte Carlo generated spectra and a new descriptive parameter.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Chofor, Ndimofor; Harder, Dietrich; Willborn, Kay; Rühmann, Antje; Poppe, Björn</p> <p>2011-09-01</p> <p>The varying low-<span class="hlt">energy</span> contribution to the photon spectra at points within and around radiotherapy photon fields is associated with variations in the responses of non-water equivalent dosimeters and in the water-to-material dose conversion factors for tissues such as the red bone marrow. In addition, the presence of low-<span class="hlt">energy</span> photons in the photon spectrum enhances the RBE in general and in particular for the induction of second malignancies. The present study discusses the general rules valid for the low-<span class="hlt">energy</span> spectral component of radiotherapeutic photon beams at points within and in the periphery of the treatment field, taking as an example the Siemens Primus linear accelerator at 6 MV and 15 MV. The photon spectra at these points and their typical variations due to the target system, attenuation, single and multiple Compton scattering, are described by the Monte Carlo method, using the code BEAMnrc/EGSnrc. A survey of the role of low <span class="hlt">energy</span> photons in the spectra within and around radiotherapy fields is presented. In addition to the spectra, some data compression has proven useful to support the overview of the behaviour of the low-<span class="hlt">energy</span> component. A characteristic indicator of the presence of low-<span class="hlt">energy</span> photons is the dose fraction attributable to photons with <span class="hlt">energies</span> not exceeding 200 <span class="hlt">keV</span>, termed P(D)(200 <span class="hlt">keV</span>). Its values are calculated for different depths and lateral positions within a water phantom. For a pencil beam of 6 or 15 MV primary photons in water, the radial distribution of P(D)(200 <span class="hlt">keV</span>) is bellshaped, with a wide-ranging exponential tail of half value 6 to 7 cm. The P(D)(200 <span class="hlt">keV</span>) value obtained on the central axis of a photon field shows an approximately proportional increase with field size. Out-of-field P(D)(200 <span class="hlt">keV</span>) values are up to an order of magnitude higher than on the central axis for the same irradiation depth. The 2D pattern of P(D)(200 <span class="hlt">keV</span>) for a radiotherapy field visualizes the regions, e.g. at the field margin, where changes of</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014NIMPB.336..135K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014NIMPB.336..135K"><span>Teflon impregnated anatase TiO2 nanoparticles irradiated by 80 <span class="hlt">keV</span> Xe+ ions</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Khanam, Rizwin; Paul, Nibedita; Kumar, P.; Kanjilal, D.; Ahmed, Gazi A.; Mohanta, Dambarudhar</p> <p>2014-10-01</p> <p>We report the effect of 80 <span class="hlt">keV</span> Xe+ ion irradiation on the morphological and optical responses of TiO2 nanoparticles spread over commercially available polytetrafluoroethylene (PTFE, Teflon). These nanoparticles were synthesized via a convenient, sol-gel approach with titanium isopropoxide as the main precursor. From X-ray diffraction (XRD) studies we found that, the nanoparticles crystallize in anatase phase and with a preferential orientation of crystallites along (1 0 1) plane. Upon irradiation at a fluence of 1.25 × 1017 ions/cm2, the nanoparticle dimension was found to increase from a value of ∼9 nm to ∼20-30 nm. Essentially, particle growth is predicted as a consequence of swelling behavior accompanied by the formation of Xe van der Waal crystals in isolated regions of nano-titania. Evidence of nanoripples was also witnessed on the surface of the irradiated nano-titania. The morphological evolution was assessed both by atomic force and transmission electron microscopies (AFM and TEM) independently. From the UV-Vis optical absorption studies, the estimated optical <span class="hlt">band</span> gap was found to drop with increasing fluence, while refractive index exhibited a remarkable improvement. Photoluminescence (PL) studies have revealed that, the <span class="hlt">band</span> edge emission and those due to the self trapped excitons (STE) and other oxygen vacancy related ones were manifested considerably as a result of Xe ion irradiation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22649227-su-novel-kernel-based-dose-engine-kev-photon-beams','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22649227-su-novel-kernel-based-dose-engine-kev-photon-beams"><span>SU-F-T-672: A Novel Kernel-Based Dose Engine for <span class="hlt">KeV</span> Photon 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>Reinhart, M; Fast, M F; Nill, S</p> <p>2016-06-15</p> <p>Purpose: Mimicking state-of-the-art patient radiotherapy with high precision irradiators for small animals allows advanced dose-effect studies and radiobiological investigations. One example is the implementation of pre-clinical IMRT-like irradiations, which requires the development of inverse planning for <span class="hlt">keV</span> photon beams. As a first step, we present a novel kernel-based dose calculation engine for <span class="hlt">keV</span> x-rays with explicit consideration of <span class="hlt">energy</span> and material dependencies. Methods: We follow a superposition-convolution approach adapted to <span class="hlt">keV</span> x-rays, based on previously published work on micro-beam therapy. In small animal radiotherapy, we assume local <span class="hlt">energy</span> deposition at the photon interaction point, since the electron ranges in tissuemore » are of the same order of magnitude as the voxel size. This allows us to use photon-only kernel sets generated by MC simulations, which are pre-calculated for six <span class="hlt">energy</span> windows and ten base materials. We validate our stand-alone dose engine against Geant4 MC simulations for various beam configurations in water, slab phantoms with bone and lung inserts, and on a mouse CT with (0.275mm)3 voxels. Results: We observe good agreement for all cases. For field sizes of 1mm{sup 2} to 1cm{sup 2} in water, the depth dose curves agree within 1% (mean), with the largest deviations in the first voxel (4%) and at depths>5cm (<2.5%). The out-of-field doses at 1cm depth agree within 8% (mean) for all but the smallest field size. In slab geometries, the mean agreement was within 3%, with maximum deviations of 8% at water-bone interfaces. The γ-index (1mm/1%) passing rate for a single-field mouse irradiation is 71%. Conclusion: The presented dose engine yields an accurate representation of keV-photon doses suitable for inverse treatment planning for IMRT. It has the potential to become a significantly faster yet sufficiently accurate alternative to full MC simulations. Further investigations will focus on <span class="hlt">energy</span> sampling as well as</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20170003140&hterms=energy&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Denergy','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20170003140&hterms=energy&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Denergy"><span>Electron Flux Models for Different <span class="hlt">Energies</span> at Geostationary Orbit</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Boynton, R. J.; Balikhin, M. A.; Sibeck, D. G.; Walker, S. N.; Billings, S. A.; Ganushkina, N.</p> <p>2016-01-01</p> <p>Forecast models were derived for energetic electrons at all <span class="hlt">energy</span> ranges sampled by the third-generation Geostationary Operational Environmental Satellites (GOES). These models were based on Multi-Input Single-Output Nonlinear Autoregressive Moving Average with Exogenous inputs methodologies. The model inputs include the solar wind velocity, density and pressure, the fraction of time that the interplanetary magnetic field (IMF) was southward, the IMF contribution of a solar wind-magnetosphere coupling function proposed by Boynton et al. (2011b), and the Dst index. As such, this study has deduced five new 1 h resolution models for the low-<span class="hlt">energy</span> electrons measured by GOES (30-50 <span class="hlt">keV</span>, 50-100 <span class="hlt">keV</span>, 100-200 <span class="hlt">keV</span>, 200-350 <span class="hlt">keV</span>, and 350-600 <span class="hlt">keV</span>) and extended the existing >800 <span class="hlt">keV</span> and >2 MeV Geostationary Earth Orbit electron fluxes models to forecast at a 1 h resolution. All of these models were shown to provide accurate forecasts, with prediction efficiencies ranging between 66.9% and 82.3%.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2008PhyB..403.4309V','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2008PhyB..403.4309V"><span>Compton scattering study of electron momentum distribution in lithium fluoride using 662 <span class="hlt">keV</span> gamma radiations</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Vijayakumar, R.; Shivaramu; Ramamurthy, N.; Ford, M. J.</p> <p>2008-12-01</p> <p>Here we report the first ever 137Cs Compton spectroscopy study of lithium fluoride. The spherical average Compton profiles of lithium fluoride are deduced from Compton scattering measurements on poly crystalline sample at gamma ray <span class="hlt">energy</span> of 662 <span class="hlt">keV</span>. To compare the experimental data, we have computed the spherical average Compton profiles using self-consistent Hartree-Fock wave functions employed on linear combination of atomic orbital (HF-LCAO) approximation. The directional Compton profiles and their anisotropic effects are also calculated using the same HF-LCAO approximation. The experimental spherical average profiles are found to be in good agreement with the corresponding HF-LCAO calculations and in qualitative agreement with Hartree-Fock free atom values. The present experimental isotropic and calculated directional profiles are also compared with the available experimental isotropic and directional Compton profiles using 59.54 and 159 <span class="hlt">keV</span> γ-rays.</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('https://www.osti.gov/biblio/22391635-mass-attenuation-coefficient-binderless-pre-treated-tannin-based-rhizophora-spp-particleboards-using-kev-photon-energy-range','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22391635-mass-attenuation-coefficient-binderless-pre-treated-tannin-based-rhizophora-spp-particleboards-using-kev-photon-energy-range"><span>Mass attenuation coefficient of binderless, pre-treated and tannin-based Rhizophora spp. particleboards using 16.59 – 25.26 <span class="hlt">keV</span> photon <span class="hlt">energy</span> range</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>Mohd Yusof, Mohd Fahmi, E-mail: mfahmi@usm.my; Hamid, Puteri Nor Khatijah Abdul; Tajuddin, Abdul Aziz</p> <p>2015-04-29</p> <p>The Rhizophora spp. particleboards were fabricated using ≤ 104 µm particle size at three different fabrication methods; binderless, steam pre-treated and tannin-added. The mass attenuation coefficient of Rhizophora spp. particleboards were measured using x-ray fluorescent (XRF) photon from niobium, molybdenum, palladium, silver and tin metal plates that provided photon <span class="hlt">energy</span> between 16.59 to 25.26 <span class="hlt">keV</span>. The results were compared to theoretical values for water calculated using photon cross-section database (XCOM).The results showed that all Rhizophora spp. particleboards having mass attenuation coefficient close to calculated XCOM for water. Tannin-added Rizophora spp. particleboard was nearest to calculated XCOM for water with χ2 valuemore » of 13.008 followed by binderless Rizophora spp. (25.859) and pre-treated Rizophora spp. (91.941)« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20060013027','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20060013027"><span>INTEGRAL Observations of the Galactic 511 <span class="hlt">keV</span> Emission and MeV Gamma-ray Astrophysics</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Watanabe, Ken</p> <p>2005-01-01</p> <p>Although there are a number of interesting phenomena, such as Nucleosynthesis in stars, in the MeV <span class="hlt">energy</span> region, the observations have been difficult due to a small signal to noise (background) ratio (less than 1%). While NASA's Compton Gamma-ray Observatory (CGRO) enabled us to explore the Gamma-ray universe, ESA's INTEGRAL mission, launched in 2002, is providing us more detailed information with its superior <span class="hlt">energy</span> and angular resolution. We will briefly discuss some of the current issues in MeV Gamma-ray Astrophysics. Then, we will focus on the Galactic 511 <span class="hlt">keV</span> emission with the latest INTEGRAL observations, and talk about challenges we currently have.</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('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 additional 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/2018JInst..13C4002S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JInst..13C4002S"><span>VOXES: a high precision X-ray spectrometer for diffused sources with HAPG crystals in the 2–20 <span class="hlt">keV</span> range</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Scordo, A.; Curceanu, C.; Miliucci, M.; Shi, H.; Sirghi, F.; Zmeskal, J.</p> <p>2018-04-01</p> <p>Bragg spectroscopy is one of the best established experimental methods for high <span class="hlt">energy</span> resolution X-ray measurements and has been widely used in several fields, going from fundamental physics to quantum mechanics tests, synchrotron radiation and X-FEL applications, astronomy, medicine and industry. However, this technique is limited to the measurement of photons produced from well collimated or point-like sources and becomes quite inefficient for photons coming from extended and diffused sources like those, for example, emitted in the exotic atoms radiative transitions. The VOXES project's goal is to realise a prototype of a high resolution and high precision X-ray spectrometer, using Highly Annealed Pyrolitic Graphite (HAPG) crystals in the Von Hamos configuration, working also for extended sources. The aim is to deliver a cost effective system having an <span class="hlt">energy</span> resolution at the level of eV for X-ray <span class="hlt">energies</span> from about 2 <span class="hlt">keV</span> up to tens of <span class="hlt">keV</span>, able to perform sub-eV precision measurements with non point-like sources. In this paper, the working principle of VOXES, together with first results, are presented.</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('http://adsabs.harvard.edu/abs/2017PPN....48..691B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017PPN....48..691B"><span>Low-<span class="hlt">energy</span> nuclear reactions in crystal structures</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bagulya, A. V.; Dalkarov, O. D.; Negodaev, M. A.; Rusetskii, A. S.</p> <p>2017-09-01</p> <p>Results of studying low-<span class="hlt">energy</span> nuclear reactions at the HELIS facility (LPI) are presented. Investigations of yields from DD reactions in deuterated crystal structures at deuteron <span class="hlt">energies</span> of 10 to 25 <span class="hlt">keV</span> show a considerable enhancement effect. It is shown that exposure of the deuterated targets to the H+ (proton) and Ne+ beams with <span class="hlt">energies</span> from 10 to 25 <span class="hlt">keV</span> and an X-ray beam with the <span class="hlt">energy</span> of 20 to 30 <span class="hlt">keV</span> stimulates DD reaction yields. For the CVD diamond target, it is shown that its orientation with respect to the deuteron beam affects the neutron yield. The D+ beam is shown to cause much higher heat release in the TiDx target than the H+ and Ne+ beams, and this heat release depends on the deuterium concentration in the target and the current density of the deuteron beam.</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 addition, 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('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('https://ntrs.nasa.gov/search.jsp?R=20030020650&hterms=Nuclear+explosion&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3DNuclear%2Bexplosion','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20030020650&hterms=Nuclear+explosion&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3DNuclear%2Bexplosion"><span>B-MINE: The Balloon-Borne Microcalorimeter Nuclear Line Explorer</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Silver, E.; Schnopper, H.; Jones, C.; Forman, W.; Bandler, S.; Murray, S.; Romaine, S.; Slane, P.; Grindlay, J.; Madden, N.</p> <p>2001-01-01</p> <p>B-MINE is a concept for a balloon mission designed to probe the deepest regions of a supernova explosion by detecting 44Ti emission at 68 <span class="hlt">keV</span> with spatial and spectral resolutions that are sufficient to determine the extent and velocity distribution of the 44Ti emitting region. The payload introduces the concept of focusing optics and microcalorimeter spectroscopy to nuclear line emission astrophysics. B-MINE has a thin, plastic foil telescope multilayered to maximize the reflectivity in a 20 <span class="hlt">keV</span> <span class="hlt">band</span> centered at 68 <span class="hlt">keV</span> and a microcalorimeter array optimized for the same <span class="hlt">energy</span> <span class="hlt">band</span>. This combination provides a reduced background, an <span class="hlt">energy</span> resolution of 50 eV and a 3 sigma sensitivity in 10 (exp 6) s of 3.3 x 10(exp -7) ph cm(exp -2) s(exp -1) at 68 <span class="hlt">keV</span>. During the course of a long duration balloon flight, B-MINE could carry out a detailed study of the 44Ti emission line centroid and width in CAS A.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017PhRvD..96j3007G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017PhRvD..96j3007G"><span>Measurement of light and charge yield of low-<span class="hlt">energy</span> electronic recoils in liquid xenon</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Goetzke, L. W.; Aprile, E.; Anthony, M.; Plante, G.; Weber, M.</p> <p>2017-11-01</p> <p>The dependence of the light and charge yield of liquid xenon on the applied electric field and recoil <span class="hlt">energy</span> is important for dark matter detectors using liquid xenon time projections chambers. Few measurements have been made of this field dependence at recoil <span class="hlt">energies</span> less than 10 <span class="hlt">keV</span>. In this paper, we present results of such measurements using a specialized detector. Recoil <span class="hlt">energies</span> are determined via the Compton coincidence technique at four drift fields relevant for liquid xenon dark matter detectors: 0.19, 0.48, 1.02, and 2.32 kV /cm . Mean recoil <span class="hlt">energies</span> down to 1 <span class="hlt">keV</span> were measured with unprecedented precision. We find that the charge and light yield are anticorrelated above ˜3 <span class="hlt">keV</span> and that the field dependence becomes negligible below ˜6 <span class="hlt">keV</span> . However, below 3 <span class="hlt">keV</span>, we find a charge yield significantly higher than expectation and a reconstructed <span class="hlt">energy</span> deviating from linearity.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24144616','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24144616"><span>Calculations of stopping powers and inelastic mean free paths for 20 eV-20 <span class="hlt">keV</span> electrons in 11 types of human tissue.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Tan, Zhenyu; Liu, Wei</p> <p>2013-12-01</p> <p>Systematic calculations are performed for determining the stopping powers (SP) and inelastic mean free paths (IMFP) for 20 eV-20 <span class="hlt">keV</span> electrons in 11 types of human tissue. The calculations are based on a dielectric model, including the Born-Ochkur exchange correction. The optical <span class="hlt">energy</span> loss functions (OELF) are empirically evaluated, because of the lack of available experimental optical data for the 11 tissues under consideration. The evaluated OELFs are examined by the f-sum rule expected from the dielectric response theory, and by calculation of the mean excitation <span class="hlt">energy</span>. The calculated SPs are compared with those for PMMA (polymethylmethacrylate, a tissue equivalent material) and liquid water. The SP and IMFP data presented here are the results for the 11 human tissues over the <span class="hlt">energy</span> range of 20 eV-20 <span class="hlt">keV</span>, and are of importance in radiotherapy planning and for studies of various radiation effects on human tissues. © 2013 Elsevier Ltd. All rights reserved.</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/1981PhRvC..24.1961H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1981PhRvC..24.1961H"><span>206Pb+n resonances for E=600-900 <span class="hlt">keV</span>: Neutron strength functions</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Horen, D. J.; Harvey, J. A.; Hill, N. W.</p> <p>1981-11-01</p> <p>Data from high resolution neutron transmission and differential scattering measurements performed on 206Pb have been analyzed for E=600-900 <span class="hlt">keV</span>. Resonance parameters (i.e., E, l, J, and Γn) have been deduced for many of the 161 resonances observed. Strength functions and potential phase shifts for s-, p-, and d-wave neutrons for En-0-900 <span class="hlt">keV</span> are compared with optical model calculations. It is found that the phase contributed by the external R function as well as the integrated neutron strength functions can be reproduced for the s and d waves with a well depth of V0=50.4 MeV for the real potential and WD=6.0 MeV for an imaginary surface potential. Somewhat smaller values (V0=48.7 MeV and WD=2.0 MeV) are required to reproduce the p-wave data. These values of the real potential are also found to give the experimentally observed binding <span class="hlt">energies</span> for the 4s12, 3d32, and 3d52 single particle levels (V0=50.4 MeV), and the 3p12 single particle level (V0=48.7 MeV). Nuclear level densities for s and d waves are found to be well represented by a constant temperature model. However, the model under estimates the number of p-wave resonances. NUCLEAR REACTIONS 206Pb(n), (n,n), E=600-900 <span class="hlt">keV</span>; measured σT(E), σ(E,θ). 207Pb deduced resonance parameters, Jπ, Γn, neutron strength functions, optical model parameters for l=0,1,2.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28495460','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28495460"><span>Dual-<span class="hlt">energy</span> computed tomography for the detection of focal liver lesions.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Lago, K N; Vallejos, J; Capuñay, C; Salas, E; Reynoso, E; Carpio, J B; Carrascosa, P M</p> <p></p> <p>To qualitatively and quantitatively explore the spectral study of focal liver lesions, comparing it with the usual polychromatic assessment with single-<span class="hlt">energy</span> computed tomography. We prospectively studied 50 patients with at least one focal liver lesion who were referred for abdominal multidetector computed tomography with intravenous contrast material. The portal phase was acquired with dual <span class="hlt">energy</span> sources. The density of the lesions and of the surrounding liver parenchyma was measured both in the baseline polychromatic acquisition and in the posterior monochromatic reconstructions at 40 <span class="hlt">keV</span>, 70 <span class="hlt">keV</span>, and 140 <span class="hlt">keV</span>. Spectral curves were traced and the dual-<span class="hlt">energy</span> indices and contrast-to-noise ratio were calculated. Lastly, the quality of the images and the detectability of the lesions were assessed qualitatively. Densitometric differences between the different types of lesions (avascular and vascularized) and the liver were greater at low <span class="hlt">energy</span> levels (left side of the spectral curve) than in the polychromatic evaluation. In the subjective assessment, the 40<span class="hlt">keV</span> <span class="hlt">energy</span> level had the greatest lesion detectability. Monochromatic spectral study with dual-<span class="hlt">energy</span> computed tomography provides better lesion detectability at 40<span class="hlt">keV</span> compared to that provided by the ordinary polychromatic evaluation. Copyright © 2017 SERAM. Publicado por Elsevier España, S.L.U. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2001AIPC..576..900S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2001AIPC..576..900S"><span>Recoil implantation of boron into silicon by high <span class="hlt">energy</span> silicon ions</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Shao, L.; Lu, X. M.; Wang, X. M.; Rusakova, I.; Mount, G.; Zhang, L. H.; Liu, J. R.; Chu, Wei-Kan</p> <p>2001-07-01</p> <p>A recoil implantation technique for shallow junction formation was investigated. After e-gun deposition of a B layer onto Si, 10, 50, or 500 <span class="hlt">keV</span> Si ion beams were used to introduce surface deposited B atoms into Si by knock-on. It has been shown that recoil implantation with high <span class="hlt">energy</span> incident ions like 500 <span class="hlt">keV</span> produces a shallower B profile than lower <span class="hlt">energy</span> implantation such as 10 <span class="hlt">keV</span> and 50 <span class="hlt">keV</span>. This is due to the fact that recoil probability at a given angle is a strong function of the <span class="hlt">energy</span> of the primary projectile. Boron diffusion was showed to be suppressed in high <span class="hlt">energy</span> recoil implantation and such suppression became more obvious at higher Si doses. It was suggested that vacancy rich region due to defect imbalance plays the role to suppress B diffusion. Sub-100 nm junction can be formed by this technique with the advantage of high throughput of high <span class="hlt">energy</span> implanters.</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('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/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. Additionally, 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/2015APS..DPPTP2003M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015APS..DPPTP2003M"><span>Understanding Bright 13 <span class="hlt">keV</span> Kr K-shell X-ray Sources at the NIF</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>May, M. J.; Colvin, J. D.; Kemp, G. E.; Fournier, K. B.; Scott, H.; Patel, M.; Barrios, Widmann; Widmann, K.</p> <p>2015-11-01</p> <p>High x-ray conversion efficiency (CE) K-shell Kr sources are being investigated for High <span class="hlt">Energy</span> Density experiments. These sources are 4.1 mm in diameter 4.4 mm tall hollow epoxy tubes having a 40 μm thick wall holding either 1.2 or 1.5 atm of Kr gas. The CE of K-shell Kr is dependent upon the peak electron temperature in the radiating plasma. In the NIF experiments, the available <span class="hlt">energy</span> heats the source to Te = 6-7 <span class="hlt">keV</span>, well below the temperature of Te ~25 <span class="hlt">keV</span> needed to optimize the Kr CE. The CE is a steep function of the peak electron temperature. A spatially averaged electron temperature can be estimated from measured He(α) and Ly(α) line ratios. Some disagreement has been observed in the simulated and measured line ratios for some of these K-shell sources. Disagreements have been observed between the simulated and measured line ratios for some of these K-shell sources. To help understand this issue, Kr gas pipes have been shot with 3 ω light at ?750 kJ at ~210, ~140 TW and ~120 TW power levels with 3.7, 5.2 and 6.7 ns pulses, respectively. The power and pulse length scaling of the measured CE and K-shell line ratios and their comparison to simulations will be discussed. This work was performed under the auspic</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> </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('http://adsabs.harvard.edu/abs/2007ApJS..172..341C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2007ApJS..172..341C"><span>The XMM-Newton Wide-Field Survey in the COSMOS Field. II. X-Ray Data and the logN-logS Relations</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Cappelluti, N.; Hasinger, G.; Brusa, M.; Comastri, A.; Zamorani, G.; Böhringer, H.; Brunner, H.; Civano, F.; Finoguenov, A.; Fiore, F.; Gilli, R.; Griffiths, R. E.; Mainieri, V.; Matute, I.; Miyaji, T.; Silverman, J.</p> <p>2007-09-01</p> <p>We present data analysis and X-ray source counts for the first season of XMM-Newton observations in the COSMOS field. The survey covers ~2 deg2 within the region of sky bounded by 09h57m30s<R.A.<10h03m30s, 01deg27'30''<decl.<02deg57'30'' with a total net integration time of 504 ks. A maximum likelihood source detection was performed in the 0.5-2, 2-4.5, and 4.5-10 <span class="hlt">keV</span> <span class="hlt">energy</span> <span class="hlt">bands</span>, and 1390 pointlike sources were detected in at least one <span class="hlt">band</span>. Detailed Monte Carlo simulations were performed to fully test the source-detection method and to derive the sky coverage to be used in the computation of the logN-logS relations. These relations were then derived in the 0.5-2, 2-10, and 5-10 <span class="hlt">keV</span> <span class="hlt">energy</span> <span class="hlt">bands</span>, down to flux limits of 7.2×10-16 ergs cm-2 s-1, 4.0×10-15 ergs cm-2 s-1, and 9.7×10-15 ergs cm-2 s-1, respectively. Thanks to the large number of sources detected in the COSMOS survey, the logN-logS curves are tightly constrained over a range of fluxes which were poorly covered by previous surveys, especially in the 2-10 and 5-10 <span class="hlt">keV</span> <span class="hlt">bands</span>. The 0.5-2 and 2-10 <span class="hlt">keV</span> differential logN-logS relations were fitted with a broken power-law model which revealed a Euclidean slope at the bright end and a flatter slope (α~1.5) at faint fluxes. In the 5-10 <span class="hlt">keV</span> <span class="hlt">energy</span> <span class="hlt">band</span> a single power law provides an acceptable fit to the observed source counts with a slope α~2.4. A comparison with the results of previous surveys shows good agreement in all the <span class="hlt">energy</span> <span class="hlt">bands</span> under investigation in the overlapping flux range. We also notice a remarkable agreement between our logN-logS relations and the most recent model of the X-ray background. Based on observations obtained with XMM-Newton, an ESA science mission with instruments and contributions directly funded by ESA Member States and NASA; also based on data collected at the Canada-France-Hawaii Telescope operated by the National Research Council of Canada, the Centre National de la Recherche Scientifique de France, and the University of</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20050166901&hterms=nature&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D90%26Ntt%3Dnature','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20050166901&hterms=nature&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D90%26Ntt%3Dnature"><span>The Physical Nature of the Sharp Spectral Feature at 7 <span class="hlt">keV</span> Detected in 1H0707-495</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Brandt, Niel</p> <p>2005-01-01</p> <p>XMM-Newton acquired data on the accepted target, 1H0707-495, on 2002 October 13 during revolution 0521. The observation was successful, with only about 5% data loss due to background flaring. We compared the data from this observation with earlier data taken on this Narrow-Line Seyfert 1 about two years before, performing interpretation studies in the context of the partial-covering model. Our second longer observation once again displays a sharp (< 200 eV) spectral drop above 7 <span class="hlt">keV</span>. However, in comparison to the first observation, the edge depth and <span class="hlt">energy</span> have changed significantly. In addition to changes in the edge parameters, the high-<span class="hlt">energy</span> spectrum appears steeper. The changes in the high-<span class="hlt">energy</span> spectrum can be adequately explained in terms of a partial-covering absorber out-flowing from the central region. The low-<span class="hlt">energy</span> spectrum also shows significant long-term spectral variability, including (1) a substantial increase in the disk temperature, (2) detection of an approx. 0.9 <span class="hlt">keV</span> emission feature, and (3) the presence of ionized absorption that was detected during the ASCA mission. The large increase in disk temperature, and the more modest rise in luminosity, can be understood if we consider a slim-disk model for 1H0707-495. In addition, the higher disk luminosity could be the driving force behind the outflow and the re-appearance of an ionized medium during the second XMM-Newton observation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26524308','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26524308"><span>X-ray spectroscopy for chemistry in the 2-4 <span class="hlt">keV</span> <span class="hlt">energy</span> regime at the XMaS beamline: ionic liquids, Rh and Pd catalysts in gas and liquid environments, and Cl contamination in γ-Al2O3.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Thompson, Paul B J; Nguyen, Bao N; Nicholls, Rachel; Bourne, Richard A; Brazier, John B; Lovelock, Kevin R J; Brown, Simon D; Wermeille, Didier; Bikondoa, Oier; Lucas, Christopher A; Hase, Thomas P A; Newton, Mark A</p> <p>2015-11-01</p> <p>The 2-4 <span class="hlt">keV</span> <span class="hlt">energy</span> range provides a rich window into many facets of materials science and chemistry. Within this window, P, S, Cl, K and Ca K-edges may be found along with the L-edges of industrially important elements from Y through to Sn. Yet, compared with those that cater for <span class="hlt">energies</span> above ca. 4-5 <span class="hlt">keV</span>, there are relatively few resources available for X-ray spectroscopy below these <span class="hlt">energies</span>. In addition, in situ or operando studies become to varying degrees more challenging than at higher X-ray <span class="hlt">energies</span> due to restrictions imposed by the lower <span class="hlt">energies</span> of the X-rays upon the design and construction of appropriate sample environments. The XMaS beamline at the ESRF has recently made efforts to extend its operational <span class="hlt">energy</span> range to include this softer end of the X-ray spectrum. In this report the resulting performance of this resource for X-ray spectroscopy is detailed with specific attention drawn to: understanding electrostatic and charge transfer effects at the S K-edge in ionic liquids; quantification of dilution limits at the Cl K- and Rh L3-edges and structural equilibria in solution; in vacuum deposition and reduction of [Rh(I)(CO)2Cl]2 to γ-Al2O3; contamination of γ-Al2O3 by Cl and its potential role in determining the chemical character of supported Rh catalysts; and the development of chlorinated Pd catalysts in `green' solvent systems. Sample environments thus far developed are also presented, characterized and their overall performance evaluated.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1422617','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1422617"><span>Effects of electronic excitation in 150 <span class="hlt">keV</span> Ni ion irradiation of metallic 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>Zarkadoula, Eva; Samolyuk, German; Weber, William J.</p> <p></p> <p>We use the two-temperature model in molecular dynamic simulations of 150 <span class="hlt">keV</span> Ni ion cascades in nickel and nickel-based alloys to investigate the effect of the <span class="hlt">energy</span> exchange between the atomic and the electronic systems during the primary stages of radiation damage. We find that the electron-phonon interactions result in a smaller amount of defects and affect the cluster formation, resulting in smaller clusters. These results indicate that ignoring the local heating due to the electrons results in the overestimation of the amount of damage and the size of the defect clusters. A comparison of the average defect production tomore » the Norgett-Robinson-Torrens (NRT) prediction over a range of <span class="hlt">energies</span> is provided.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1422617-effects-electronic-excitation-kev-ni-ion-irradiation-metallic-systems','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1422617-effects-electronic-excitation-kev-ni-ion-irradiation-metallic-systems"><span>Effects of electronic excitation in 150 <span class="hlt">keV</span> Ni ion irradiation of metallic systems</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Zarkadoula, Eva; Samolyuk, German; Weber, William J.</p> <p>2018-01-18</p> <p>We use the two-temperature model in molecular dynamic simulations of 150 <span class="hlt">keV</span> Ni ion cascades in nickel and nickel-based alloys to investigate the effect of the <span class="hlt">energy</span> exchange between the atomic and the electronic systems during the primary stages of radiation damage. We find that the electron-phonon interactions result in a smaller amount of defects and affect the cluster formation, resulting in smaller clusters. These results indicate that ignoring the local heating due to the electrons results in the overestimation of the amount of damage and the size of the defect clusters. A comparison of the average defect production tomore » the Norgett-Robinson-Torrens (NRT) prediction over a range of <span class="hlt">energies</span> is provided.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19830028765','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19830028765"><span>Evidence for variability of the hard X-ray feature in the Hercules X-1 <span class="hlt">energy</span> spectrum</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Tueller, J.; Cline, T. L.; Teegarden, B. J.; Paciesas, W. S.; Boclet, D.; Durochoux, P.; Hameury, J. M.; Prantzos, N.; Haymes, R. C.</p> <p>1983-01-01</p> <p>The hard X-ray spectrum of HER X-1 was measured for the first time with a high resolution (1.4 <span class="hlt">keV</span> FWHM) germanium spectrometer. The observation was performed near the peak of the on-state in the 35 day cycle and the 1.24 pulsations were observed between the <span class="hlt">energies</span> of 20 <span class="hlt">keV</span> and 70 <span class="hlt">keV</span>. The feature corresponds to an excess of 7.5 sigma over the low <span class="hlt">energy</span> continuum. Smooth continuum models are poor fits to the entire <span class="hlt">energy</span> range (chance probabilities of 2 percent or less). The best fit <span class="hlt">energies</span> are 35 <span class="hlt">keV</span> for an absorption line and 39 <span class="hlt">keV</span> for an emission line. These are significantly lower <span class="hlt">energies</span> than those derived from previous experiments. A direct comparison of our data with the results of the MPI/AIT group shows statistically significant variations which strongly suggest variability in the source.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1985PhRvA..32..959U','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1985PhRvA..32..959U"><span>Photoeffect cross sections of some rare-earth elements at 145.4 <span class="hlt">keV</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Umesh, T. K.; Ranganathaiah, C.; Sanjeevaiah, B.</p> <p>1985-08-01</p> <p>Total attenuation cross sections in the elements La, Ce, Pr, Nd, Sm, Gd, Dy, Ho, and Er were derived from the measured total cross sections of their simple oxide compounds, by employing the mixture rule at 145.4-<span class="hlt">keV</span> photon <span class="hlt">energy</span>. The compound cross sections have been measured by performing transmission experiments in a good geometry setup. From the derived total cross sections of elements, photoeffect cross sections have been obtained by subtracting the theoretical scattering cross sections. A good agreement is observed between the present data of photoeffect cross sections and Scofield's theoretical data.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018EPJWC.16501042P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018EPJWC.16501042P"><span>The 10B(p,α)7Be S(E)-factor from 5 <span class="hlt">keV</span> to 1.5 MeV using the Trojan Horse Method</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Puglia, Sebastiana Maria Regina; Spitaleri, Claudio; La Cognata, Marco; Lamia, Livio; Broggini, Carlo; Caciolli, Antonio; Carlin, Nelson; Cherubini, Silvio; Cvetinovic, Alexandra; D'Agata, Giuseppe; Dell'aquila, Daniele; Depalo, Rosanna; Gulino, Marisa; Guardo, Giovanni Luca; Indelicato, Iolanda; Lombardo, Ivano; Menegazzo, Roberto; Munhoz, Marcelo Gimenez; Pizzone, Rosario Gianluca; Rapisarda, Giuseppe Gabriele; Rigato, Valentino; Romano, Stefano; Sergi, Maria Letizia; Souza, Francisco; Sparta, Roberta; Tudisco, Salvo; Tumino, Aurora</p> <p>2018-01-01</p> <p>The 10B(p,α)7Be reaction is the main responsible for the 10B destruction in stellar interior [1]. In such environments this p-capture process occurs at a Gamow <span class="hlt">energy</span> of 10 <span class="hlt">keV</span> and takes places mainly through a resonant state (Ex = 8.701 MeV) of the compound 11C nucleus. Thus a resonance right in the region of the Gamow peak is expected to significantly influence the behavior of the astrophysical S(E)-factor. The 10B(p,α)7Be reaction was studied via the Trojan Horse Method (THM) applied to the 2H(10B,α7Be)n in order to extract the astrophysical S(E)-factor in a wide <span class="hlt">energy</span> range from 5 <span class="hlt">keV</span> to 1.5 MeV.</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/1999ApJ...512..920S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1999ApJ...512..920S"><span>High-<span class="hlt">Energy</span> Spectral and Temporal Characteristics of GRO J1008-57</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Shrader, C. R.; Sutaria, F. K.; Singh, K. P.; Macomb, D. J.</p> <p>1999-02-01</p> <p>A transient X-ray source, GRO J1008-57, was discovered by the Burst and Transient Source Experiment (BATSE) on board the Compton Gamma Ray Observatory (CGRO) in 1993 July. It reached a maximum intensity of about 1.4 times that of the Crab, in the 20-60 <span class="hlt">keV</span> <span class="hlt">energy</span> <span class="hlt">band</span>. Pulsations in the X-ray intensity were detected at a period of 93.5 s. It has subsequently been determined to be a member of the Be star subclass of X-ray transients. In addition to BATSE, GRO J1008-57 was observed during its outburst by several pointed high-<span class="hlt">energy</span> experiments: ROSAT, ASCA, and CGRO/OSSE. These nonsimultaneous but contemporaneous observations took place near and shortly after the peak of the outburst light curve. We report for the first time on a combined analysis of the CGRO and ASCA data sets. We have attempted to model the broadband high-<span class="hlt">energy</span> continuum distribution and phase-resolved spectra. The broadband, phase-averaged continuum is well approximated by a power law with an exponential cutoff. Evidence for 6.4 <span class="hlt">keV</span> line emission due to Fe is presented based on our spectral analysis. The <span class="hlt">energy</span> dependence of the pulse profiles is examined in order to determine the <span class="hlt">energy</span> at which the low-<span class="hlt">energy</span> double-peaked profile detected by ASCA evolves into single-peaked pulse profile detected by BATSE. We discuss the implications of this pulse profile for the magnetic field and beam distribution for GRO J1008-57. Analysis of the BATSE and Rossi X-Ray Timing Explorer/ASM flux histories suggests that Porbital~135 days. We further suggest that a transient disk is likely to form during episodes of outbursts.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/21137228-design-development-gamma-ray-burst-monitor-calet-mission','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/21137228-design-development-gamma-ray-burst-monitor-calet-mission"><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>Nakahira, S.; Yamaoka, K.; Yoshida, A.</p> <p></p> <p>We propose to provide a gamma-ray burst monitor (GBM) for the CALET mission to monitor gamma-ray bursts (GRBs) together with the CALET Imaging Calorimeter detector. The major purpose is to derive a wide-<span class="hlt">band</span> <span class="hlt">energy</span> spectrum of GRBs over an unprecedented 9 decades of <span class="hlt">energy</span> (from a few <span class="hlt">keV</span> to a few TeV) in combination with the CALET tower detector. Hence it is desirable to have the CALET-GBM covering an <span class="hlt">energy</span> range from a few <span class="hlt">keV</span> to about 20 MeV to avoid a gap in the observational <span class="hlt">energy</span> <span class="hlt">bands</span>. The design of GBM is underway to fulfill this requirement. The currentmore » detector candidate is BGO, and a LaBr (Ce) scintillator which has a superior <span class="hlt">energy</span> resolution to that of NaI(Tl). In this paper, design and expected performance of the CALET-GBM is shown.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011ITNS...58..687C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011ITNS...58..687C"><span>Monitoring <span class="hlt">Energy</span> Calibration Drift Using the Scintillator Background Radiation</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Conti, Maurizio; Eriksson, Lars; Hayden, Charles</p> <p>2011-06-01</p> <p>Scintillating materials commonly used in nuclear medicine can contain traces of isotopes that naturally emit gamma or beta radiation. Examples of these are 138La contained in LaBr3 and other Lanthanum based scintillators, and 176Lu contained in LSO, LYSO, LuYAP and other Lutetium based scintillators. In particular,176Lu decays into 176Hf and emits a beta particle with maximum <span class="hlt">energy</span> 589 <span class="hlt">keV</span>, and a cascade of gamma rays of <span class="hlt">energies</span> 307 <span class="hlt">keV</span>, 202 <span class="hlt">keV</span> and 88 <span class="hlt">keV</span>. We propose to use the background radiation for monitoring of detector calibration drift and for self-calibration of detectors in complex detector systems. A calibration drift due to random or systematic changes in photomultiplier tube (PMT) gain was studied in a Siemens PET scanner, based on LSO blocks. Both a conventional radioactive source (68Ge, 511 <span class="hlt">keV</span> photons from electron-positron annihilation) and the LSO background radiation were used for calibration. The difference in the calibration peak shift at 511 <span class="hlt">keV</span> estimated with the two methods was less than 10%.</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/1980JaJAP..19.1627I','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1980JaJAP..19.1627I"><span>Origin of Si(LMM) Auger Electron Emission from Silicon and Si-Alloys by <span class="hlt">keV</span> Ar+ Ion Bombardment</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Iwami, Motohiro; Kim, Su Chol; Kataoka, Yoshihide; Imura, Takeshi; Hiraki, Akio; Fujimoto, Fuminori</p> <p>1980-09-01</p> <p>Si(LMM) Auger electrons emitted from specimens of pure silicon and several Si-alloys (Ni-Si, Pd-Si and Cu-Si) under <span class="hlt">keV</span> Ar+ ion bombardment, were examined. In the Auger spectra from all specimens studied there were four peaks at <span class="hlt">energies</span> of 92, 86, 76 and 66 eV. The Auger signal intensity varied considerably with both the incident angle and the <span class="hlt">energy</span> of the primary ion beam. It is proposed that the Auger electrons are emitted from silicon atoms (or ions) just beneath the specimen surface but free from the bulk network.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20170003471&hterms=1094&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3D%2526%25231094','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20170003471&hterms=1094&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3D%2526%25231094"><span>Atomic Scattering Factor of the ASTRO-H (Hitomi) SXT Reflector Around the Gold's L Edges</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Kikuchi, Naomichi; Kurashima, Sho; Ishida, Manabu; Iizuka, Ryo; Maeda, Yoshitomo; Hayashi, Takayuki; Okajima, Takashi; Matsumoto, Hironori; Mitsubishi, Ikuyuki; Saji, Shigetaka</p> <p>2016-01-01</p> <p>The atomic scattering factor in the <span class="hlt">energy</span> range of 11.2 - 15.4 <span class="hlt">keV</span> for the ASTRO-H Soft X-ray Telescope (SXT) is reported. The large effective area of the SXT makes use of photon spectra above 10 <span class="hlt">keV</span> viable, unlike most other X-ray satellites with total-reflection mirror optics. Presence of gold's L-edges in the <span class="hlt">energy</span> <span class="hlt">band</span> is a major issue, as it complicates the function of the effective area. In order to model the area, the reflectivity measurements in the 11.2 - 15.4 <span class="hlt">keV</span> <span class="hlt">band</span> with the <span class="hlt">energy</span> pitch of 0.4 - 0.7 eV were made in the synchrotron beam-line Spring-8 BL01B1. We obtained atomic scattering factors f1 and f2 by the curve fitting to the reflectivities of our witness sample. The edges associated with the L-I, II, and III transitions are identified, of which the depths are found to be roughly 60 shallower than those expected from the Henkes atomic scattering factor.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22663323-steep-decay-phase-shaped-curvature-effect-ii-spectral-evolution','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22663323-steep-decay-phase-shaped-curvature-effect-ii-spectral-evolution"><span>Steep Decay Phase Shaped by the Curvature Effect. II. Spectral Evolution</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>Lin, Da-Bin; Mu, Hui-Jun; Lu, Rui-Jing</p> <p></p> <p>We derive a simple analytical formula to describe the evolution of spectral index β in the steep decay phase shaped by the curvature effect with the assumption that the spectral parameters and Lorentz factor of the jet shell are the same for different latitudes. Here, the value of β is estimated in the 0.3−10 <span class="hlt">keV</span> <span class="hlt">energy</span> <span class="hlt">band</span>. For a spherical thin shell with a cutoff power-law (CPL) intrinsic radiation spectrum, the spectral evolution can be read as a linear function of observer time. For the situation with the <span class="hlt">Band</span> function intrinsic radiation spectrum, the spectral evolution may be complex. Ifmore » the observed break <span class="hlt">energy</span> of the radiation spectrum is larger than 10 <span class="hlt">keV</span>, the spectral evolution is the same as that shaped by jet shells with a CPL spectrum. If the observed break <span class="hlt">energy</span> is less than 0.3 <span class="hlt">keV</span>, the value of β would be a constant. For others, the spectral evolution can be approximated as a logarithmal function of the observer time in general.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017ATel10690....1L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017ATel10690....1L"><span>Swift/BAT Detects Increase in Hard X-ray Emission from the Ultra-compact X-ray Binary 4U 1543-624</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ludlam, Renee; Miller, Jon M.; Miller-Jones, James; Reynolds, Mark</p> <p>2017-08-01</p> <p>The Swift/BAT detected an increase in hard X-ray emission (15-50 <span class="hlt">keV</span>) coming from the ultra-compact X-ray binary 4U 1543-624 around 2017 August 9. The MAXI daily monitoring also shows a gradual increase in 2.0-20.0 <span class="hlt">keV</span> X-ray intensity as of 2017 August 19. Swift/XRT ToO monitoring of the source was triggered and shows an increase in unabsorbed flux to 1.06E-9 ergs/cm2/s in the 0.3-10.0 <span class="hlt">keV</span> <span class="hlt">energy</span> <span class="hlt">band</span> as of 2017 August 26. ATCA performed ToO observations for approximately 4 hours in the 5.5 GHz and 9.0 GHz <span class="hlt">bands</span> while the antennas were in the 1.5A array configuration from 11:25-16:09 UTC on 2017 August 23. The source was not detected in either <span class="hlt">band</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22409724-we-bremsstrahlung-laser-plasma-interaction-kev-temperature-forward-dose-attenuation-factors','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22409724-we-bremsstrahlung-laser-plasma-interaction-kev-temperature-forward-dose-attenuation-factors"><span>WE-E-18A-05: Bremsstrahlung of Laser-Plasma Interaction at <span class="hlt">KeV</span> Temperature: Forward Dose and Attenuation Factors</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>Saez-Beltran, M; Fernandez Gonzalez, F</p> <p>2014-06-15</p> <p>Purpose: To obtain an analytical empirical formula for the photon dose source term in forward direction from bremsstrahlung generated from laser-plasma accelerated electron beams in aluminum solid targets, with electron-plasma temperatures in the 10–100 <span class="hlt">keV</span> <span class="hlt">energy</span> range, and to calculate transmission factors for iron, aluminum, methacrylate, lead and concrete and air, materials most commonly found in vacuum chamber labs. Methods: Bremsstrahlung fluence is calculated from the convolution of thin-target bremsstrahlung spectrum for monoenergetic electrons and the relativistic Maxwell-Juettner <span class="hlt">energy</span> distribution for the electron-plasma. Unattenuatted dose in tissue is calculated by integrating the photon spectrum with the mass-<span class="hlt">energy</span> absorption coefficient. Formore » the attenuated dose, <span class="hlt">energy</span> dependent absorption coefficient, build-up factors and finite shielding correction factors were also taken into account. For the source term we use a modified formula from Hayashi et al., and we fitted the proportionality constant from experiments with the aid of the previously calculated transmission factors. Results: The forward dose has a quadratic dependence on electron-plasma temperature: 1 joule of effective laser <span class="hlt">energy</span> transferred to the electrons at 1 m in vacuum yields 0,72 Sv per MeV squared of electron-plasma temperature. Air strongly filters the softer part of the photon spectrum and reduce the dose to one tenth in the first centimeter. Exponential higher <span class="hlt">energy</span> tail of maxwellian spectrum contributes mainly to the transmitted dose. Conclusion: A simple formula for forward photon dose from <span class="hlt">keV</span> range temperature plasma is obtained, similar to those found in kilovoltage x-rays but with higher dose per dissipated electron <span class="hlt">energy</span>, due to thin target and absence of filtration.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014ChPhB..23f5203F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014ChPhB..23f5203F"><span><span class="hlt">Energy</span> spectrum of multi-radiation of X-rays in a low <span class="hlt">energy</span> Mather-type plasma focus device</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Farzin, M. Aghamir; Reza, A. Behbahani</p> <p>2014-06-01</p> <p>The multi-radiation of X-rays was investigated with special attention to their <span class="hlt">energy</span> spectrum in a Mather-type plasma focus device (operated with argon gas). The analysis is based on the effect of anomalous resistances. To study the <span class="hlt">energy</span> spectrum, a four-channel diode X-ray spectrometer was used along with a special set of filters. The filters were suitable for detection of medium range X-rays as well as hard X-rays with <span class="hlt">energy</span> exceeding 30 <span class="hlt">keV</span>. The results indicate that the anomalous resistivity effect during the post pinch phase may cause multi-radiation of X-rays with a total duration of 300 ± 50 ns. The significant contribution of Cu—Kα was due to the medium range X-rays, nonetheless, hard X-rays with <span class="hlt">energies</span> greater than 15 <span class="hlt">keV</span> also participate in the process. The total emitted X-ray <span class="hlt">energy</span> in the forms of Cu—Kα and Cu—Kβ was around 0.14 ± 0.02 (J/Sr) and 0.04 ± 0.01 (J/Sr), respectively. The total <span class="hlt">energy</span> of the emitted hard X-ray (> 15 <span class="hlt">keV</span>) was around 0.12 ± 0.02 (J/Sr).</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> </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('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('https://www.osti.gov/biblio/22210406-capture-neutron-excited-states-sup-nucleus-taking-account-resonance-kev','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22210406-capture-neutron-excited-states-sup-nucleus-taking-account-resonance-kev"><span>Capture of a neutron to excited states of a {sup 9}Be nucleus taking into account resonance at 622 <span class="hlt">keV</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>Dubovichenko, S. B., E-mail: dubovichenko@gmail.com</p> <p>2013-10-15</p> <p>Radiative capture of a neutron to the ground and excited states of the 9Be nucleus is considered using the potential cluster model with forbidden states and with classification of cluster states by the Young schemes taking into account resonance at 622 <span class="hlt">keV</span> for thermal and astrophysical <span class="hlt">energies</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20040071148&hterms=sources+energy&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dsources%2Benergy','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20040071148&hterms=sources+energy&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dsources%2Benergy"><span>The BATSE Earth Occultation Catalog of Low <span class="hlt">Energy</span> Gamma-Ray Sources</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Harmon, B. A.; Wilson-Hodge, C. A.; Fishman, G. J.; Paciesas, W. S.; Zhang, S. N.; Finger, M. H.; Connaughton, V.; Koshut, T. M.; Henze, W.; McCollough, M. L.</p> <p>2004-01-01</p> <p>The Burst and Transient Source Experiment (BATSE),aboard the COmptOn Gamma Ray Observatory (CGRO), provided a record of the hard X-ray/low <span class="hlt">energy</span> gamma ray sky between April 1991 and June 2000. During that time, a catalog of known sources was derived from existing catalogs such as HEAO A-4 (Levine et al. 19841, as well as new transient sources discovered with RATSE and other X-ray monitors operating in the CGRO era. The Earth Occultation Technique (Harmon et al. 2001, astro-ph/0109069) was used to monitor a combination of these sources, mostly galactic, totaling about 175 objects. The catalog will present the global properties of these sources and their probability of detection (>lO mCrab, 20-100 <span class="hlt">keV</span>) with BATSE. Systematic errors due to unknown sources or background components are included. Cursory analyses to search for new transients (35-80 mCrab in the 20-100 <span class="hlt">keV</span> <span class="hlt">band</span>) and super-orbital periods in known binary sources are also presented. Whole mission light curves and associated data production/analysis tools are being delivered to the HEASARC for public use.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27410339','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27410339"><span>Highly efficient blazed grating with multilayer coating for tender X-ray <span class="hlt">energies</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Senf, F; Bijkerk, F; Eggenstein, F; Gwalt, G; Huang, Q; Kruijs, R; Kutz, O; Lemke, S; Louis, E; Mertin, M; Packe, I; Rudolph, I; Schäfers, F; Siewert, F; Sokolov, A; Sturm, J M; Waberski, Ch; Wang, Z; Wolf, J; Zeschke, T; Erko, A</p> <p>2016-06-13</p> <p>For photon <span class="hlt">energies</span> of 1 - 5 <span class="hlt">keV</span>, blazed gratings with multilayer coating are ideally suited for the suppression of stray and higher orders light in grating monochromators. We developed and characterized a blazed 2000 lines/mm grating coated with a 20 period Cr/C- multilayer. The multilayer d-spacing of 7.3 nm has been adapted to the line distance of 500 nm and the blaze angle of 0.84° in order to provide highest efficiency in the photon <span class="hlt">energy</span> range between 1.5 <span class="hlt">keV</span> and 3 <span class="hlt">keV</span>. Efficiency of the multilayer grating as well as the reflectance of a witness multilayer which were coated simultaneously have been measured. An efficiency of 35% was measured at 2 <span class="hlt">keV</span> while a maximum efficiency of 55% was achieved at 4 <span class="hlt">keV</span>. In addition, a strong suppression of higher orders was observed which makes blazed multilayer gratings a favorable dispersing element also for the low X-ray <span class="hlt">energy</span> range.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20150008983','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20150008983"><span>NuSTAR Observations of Heavily Obscured Quasars at z Is Approximately 0.5</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Lansbury, G. B.; Alexander, D. M.; Del Moro, A.; Gandhi, P.; Assef, R. J.; Stern, D.; Aird, J.; Ballantyne, D. R.; Balokovic, M.; Bauer, F. E.; <a style="text-decoration: none; " href="javascript:void(0); " onClick="displayelement('author_20150008983'); toggleEditAbsImage('author_20150008983_show'); toggleEditAbsImage('author_20150008983_hide'); "> <img style="display:inline; width:12px; height:12px; " src="images/arrow-up.gif" width="12" height="12" border="0" alt="hide" id="author_20150008983_show"> <img style="width:12px; height:12px; display:none; " src="images/arrow-down.gif" width="12" height="12" border="0" alt="hide" id="author_20150008983_hide"></p> <p>2014-01-01</p> <p>We present NuSTAR hard X-ray observations of three Type 2 quasars at z approx. = 0.4-0.5, optically selected from the Sloan Digital Sky Survey. Although the quasars show evidence for being heavily obscured, Compton-thick systems on the basis of the 2-10 <span class="hlt">keV</span> to [O(sub III)] luminosity ratio and multiwavelength diagnostics, their X-ray absorbing column densities (N(sub H)) are poorly known. In this analysis, (1) we study X-ray emission at greater than 10 <span class="hlt">keV</span>, where X-rays from the central black hole are relatively unabsorbed, in order to better constrain N(sub H). (2) We further characterize the physical properties of the sources through broad-<span class="hlt">band</span> near-UV to mid-IR spectral <span class="hlt">energy</span> distribution analyses. One of the quasars is detected with NuSTAR at greater than 8 <span class="hlt">keV</span> with a no-source probability of less than 0.1%, and its X-ray <span class="hlt">band</span> ratio suggests near Compton-thick absorption with N(sub H) is approximately greater than 5 × 10(exp 23) cm(exp -2). The other two quasars are undetected, and have low X-ray to mid-IR luminosity ratios in both the low-<span class="hlt">energy</span> (2-10 <span class="hlt">keV</span>) and high-<span class="hlt">energy</span> (10-40 <span class="hlt">keV</span>) X-ray regimes that are consistent with extreme, Compton-thick absorption (N(sub H) is approximately greater than 10(exp 24) cm(exp -2)). We find that for quasars at z is approximately 0.5, NuSTAR provides a significant improvement compared to lower <span class="hlt">energy</span> (less than 10 <span class="hlt">keV</span>) Chandra and XMM-Newton observations alone, as higher column densities can now be directly constrained.</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/2018NIMPA.882..138A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018NIMPA.882..138A"><span>Development of a 32-channel ASIC for an X-ray APD detector onboard the ISS</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Arimoto, Makoto; Harita, Shohei; Sugita, Satoshi; Yatsu, Yoichi; Kawai, Nobuyuki; Ikeda, Hirokazu; Tomida, Hiroshi; Isobe, Naoki; Ueno, Shiro; Mihara, Tatehiro; Serino, Motoko; Kohmura, Takayoshi; Sakamoto, Takanori; Yoshida, Atsumasa; Tsunemi, Hiroshi; Hatori, Satoshi; Kume, Kyo; Hasegawa, Takashi</p> <p>2018-02-01</p> <p>We report on the design and performance of a mixed-signal application specific integrated circuit (ASIC) dedicated to avalanche photodiodes (APDs) in order to detect hard X-ray emissions in a wide <span class="hlt">energy</span> <span class="hlt">band</span> onboard the International Space Station. To realize wide-<span class="hlt">band</span> detection from 20 <span class="hlt">keV</span> to 1 MeV, we use Ce:GAGG scintillators, each coupled to an APD, with low-noise front-end electronics capable of achieving a minimum <span class="hlt">energy</span> detection threshold of 20 <span class="hlt">keV</span>. The developed ASIC has the ability to read out 32-channel APD signals using 0.35 μm CMOS technology, and an analog amplifier at the input stage is designed to suppress the capacitive noise primarily arising from the large detector capacitance of the APDs. The ASIC achieves a performance of 2099 e- + 1.5 e-/pF at root mean square (RMS) with a wide 300 fC dynamic range. Coupling a reverse-type APD with a Ce:GAGG scintillator, we obtain an <span class="hlt">energy</span> resolution of 6.7% (FWHM) at 662 <span class="hlt">keV</span> and a minimum detectable <span class="hlt">energy</span> of 20 <span class="hlt">keV</span> at room temperature (20 °C). Furthermore, we examine the radiation tolerance for space applications by using a 90 MeV proton beam, confirming that the ASIC is free of single-event effects and can operate properly without serious degradation in analog and digital processing.</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/12111133','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/12111133"><span>PET attenuation coefficients from CT images: experimental evaluation of the transformation of CT into PET 511-<span class="hlt">keV</span> attenuation coefficients.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Burger, C; Goerres, G; Schoenes, S; Buck, A; Lonn, A H R; Von Schulthess, G K</p> <p>2002-07-01</p> <p>The CT data acquired in combined PET/CT studies provide a fast and essentially noiseless source for the correction of photon attenuation in PET emission data. To this end, the CT values relating to attenuation of photons in the range of 40-140 <span class="hlt">keV</span> must be transformed into linear attenuation coefficients at the PET <span class="hlt">energy</span> of 511 <span class="hlt">keV</span>. As attenuation depends on photon <span class="hlt">energy</span> and the absorbing material, an accurate theoretical relation cannot be devised. The transformation implemented in the Discovery LS PET/CT scanner (GE Medical Systems, Milwaukee, Wis.) uses a bilinear function based on the attenuation of water and cortical bone at the CT and PET <span class="hlt">energies</span>. The purpose of this study was to compare this transformation with experimental CT values and corresponding PET attenuation coefficients. In 14 patients, quantitative PET attenuation maps were calculated from germanium-68 transmission scans, and resolution-matched CT images were generated. A total of 114 volumes of interest were defined and the average PET attenuation coefficients and CT values measured. From the CT values the predicted PET attenuation coefficients were calculated using the bilinear transformation. When the transformation was based on the narrow-beam attenuation coefficient of water at 511 <span class="hlt">keV</span> (0.096 cm(-1)), the predicted attenuation coefficients were higher in soft tissue than the measured values. This bias was reduced by replacing 0.096 cm(-1) in the transformation by the linear attenuation coefficient of 0.093 cm(-1) obtained from germanium-68 transmission scans. An analysis of the corrected emission activities shows that the resulting transformation is essentially equivalent to the transmission-based attenuation correction for human tissue. For non-human material, however, it may assign inaccurate attenuation coefficients which will also affect the correction in neighbouring tissue.</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://hdl.handle.net/2060/20080036826','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20080036826"><span>ART-XC: A Medium-<span class="hlt">energy</span> X-ray Telescope System for the Spectrum-R-Gamma Mission</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Arefiev, V.; Pavlinsky, M.; Lapshov, I.; Thachenko, A.; Sazonov, S.; Revnivtsev, M.; Semena, N.; Buntov,M.; Vikhlinin, A.; Gubarev, M.; <a style="text-decoration: none; " href="javascript:void(0); " onClick="displayelement('author_20080036826'); toggleEditAbsImage('author_20080036826_show'); toggleEditAbsImage('author_20080036826_hide'); "> <img style="display:inline; width:12px; height:12px; " src="images/arrow-up.gif" width="12" height="12" border="0" alt="hide" id="author_20080036826_show"> <img style="width:12px; height:12px; display:none; " src="images/arrow-down.gif" width="12" height="12" border="0" alt="hide" id="author_20080036826_hide"></p> <p>2008-01-01</p> <p>The ART-XC instrument is an X-ray grazing-incidence telescope system in an ABRIXAS-type optical configuration optimized for the survey observational mode of the Spectrum-RG astrophysical mission which is scheduled to be launched in 2011. ART-XC has two units, each equipped with four identical X-ray multi-shell mirror modules. The optical axes of the individual mirror modules are not parallel but are separated by several degrees to permit the four modules to share a single CCD focal plane detector, 1/4 of the area each. The 450-micron-thick pnCCD (similar to the adjacent eROSITA telescope detector) will allow detection of X-ray photons up to 15 <span class="hlt">keV</span>. The field of view of the individual mirror module is about 18 x 18 arcminutes(exp 2) and the sensitivity of the ART-XC system for 4 years of survey will be better than 10(exp -12) erg s(exp -1) cm(exp -2) over the 4-12 <span class="hlt">keV</span> <span class="hlt">energy</span> <span class="hlt">band</span>. This will allow the ART-XC instrument to discover several thousand new AGNs.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20030069007&hterms=sources+energy&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dsources%2Benergy','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20030069007&hterms=sources+energy&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dsources%2Benergy"><span>The Burst and Transient Source Experiment(BATSE)Earth Occultation Catalog of Low-<span class="hlt">Energy</span> Gamma-Ray Sources. Short title: BATSE Earth Occultation Catalog v2.0</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Harmon, B. A.; Wilson, C. A.; Fishman, G. J.; Connaughton, V.; Henze, W.; Paciesas, W. S.; Finger, M. H.; McCollough, M. L.; Sahi, M.; Peterson, B.</p> <p>2003-01-01</p> <p>The Burst and Transient Source Experiment (BATSE), aboard the Compton Gamma Ray Observatory (CGRO), provided a record of the low-<span class="hlt">energy</span> gamma-ray sky (approx. 20-1000 <span class="hlt">keV</span>) between 1991 April and 2000 May (9.1y). BATSE monitored the high <span class="hlt">energy</span> sky using the Earth occultation technique (EOT) for point sources whose emission extended for times on the order of the CGRO orbital period (approx. 92m) or greater. Using the EOT to extract flux - 2 - information, a catalog of sources using data from the BATSE large area detectors has been prepared. The first part of the catalog consists of results from the all-sky monitoring of 58 sources, mostly Galactic, with intrinsic variability on timescales of hours to years. For these sources, we have included tables of flux and spectral data, and outburst times for transients. Light curves (or flux histories) covering the entire nine mission are being placed on the world wide web. We then performed a deep-sampling of these 58 objects, plus a selection of 121 more objects, combining data from the entire 9.ly BATSE dataset. Source types considered were primarily accreting binaries, but a small number of representative active galaxies, X-ray-emitting stars, and supernova remnants were also included. The sample represents a compilation of sources monitored and/or discovered with BATSE and other high <span class="hlt">energy</span> instruments between 1991 and 2000, known sources taken from the HEAO 1 A-4 (Levine et al. 1984) and Macomb and Gehrels (1999) catalogs. The deep sample results include definite detections of 82 objects and possible detections of 36 additional objects. The definite detections spanned three classes of sources: accreting black hole and neutron star binaries, active galaxies and supernova remnants. The average fluxes measured for the fourth class, the X-ray emitting stars, were below the confidence limit for definite detection. Flux data for the deep sample are presented in four <span class="hlt">energy</span> <span class="hlt">bands</span>: 20-40, 40-70, 70-160, and 160-430 <span class="hlt">keV</span>. The</p> </li> <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/2017AGUFMSH34A..03W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFMSH34A..03W"><span>Solar Wind 0.1-1 <span class="hlt">keV</span> Electrons in the Corotating Interaction Regions</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wang, L.; Tao, J.; Li, G.; Wimmer-Schweingruber, R. F.; Jian, L. K.; He, J.; Tu, C.; Tian, H.; Bale, S. D.</p> <p>2017-12-01</p> <p>Here we present a statistical study of the 0.1-1 <span class="hlt">keV</span> suprathermal electrons in the undisturbed and compressed slow/fast solar wind, for the 71 corotating interaction regions (CIRs) with good measurements from the WIND 3DP and MFI instruments from 1995 to 1997. For each of these CIRs, we separate the strahl and halo electrons based on their different behaviors in pitch angle distributions in the undisturbed and compressed solar wind. We fit both the strahl and halo <span class="hlt">energy</span> spectra to a kappa function with an index κ index and effective temperature Teff, and calculate the pitch-angle width at half-maximum (PAHM) of the strahl population. We also integrate the electron measurements between 0.1 and 1.0 <span class="hlt">keV</span> to obtain the number density n and average <span class="hlt">energy</span> Eavg for the strahl and halo populations. We find that for both the strahl and halo populations within and around these CIRs, the fitted κ index strongly correlates with Teff, similar to the quiet-time solar wind (Tao et al., ApJ, 2016). The number density of both the strahl and halo shows a strong positive correlation with the electron core temperature. The strahl number density ns is correlated with the magnitude of interplanetary magnetic field, and the strahl PAHM width is anti-correlated with the solar wind speed. These results suggest that the origin of strahl electrons from the solar corona is likely related to the electron core temperature and magnetic field strength, while the production of halo electrons in the interplanetary medium could depend on the solar wind velocity.</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('http://adsabs.harvard.edu/abs/2006APS..MARG29009P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2006APS..MARG29009P"><span>Radiation Damage From Mono-energetic Electrons Up to 200 <span class="hlt">keV</span> On Biological Systems</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Prilepskiy, Yuriy</p> <p>2006-03-01</p> <p>The electron gun of the CEBAF machine at Jefferson lab (Newport News, VA) is capable of delivering electrons with <span class="hlt">energies</span> up to 200 <span class="hlt">keV</span> with a resolution of about 10-5. This 1.5 GHz beam permits to generate cellular radiation damage within minutes. We have performed irradiation of cancer cells with different <span class="hlt">energies</span> and different currents to investigate their biological responses. This study will permit to address the physical processes involved in the RBE and LET at a level that supersedes current data listed in the literature by orders of magnitude. We will discuss the experimental setup and results of the first stage of data collected with this novel system. This research is part of a global program to provide detailed information for the understanding of radiation based cancer treatments.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1981IJFE....1...69B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1981IJFE....1...69B"><span>Design study of a 120-<span class="hlt">keV</span>, He-3 neutral beam injector</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Blum, A. S.; Barr, W. L.; Dexter, W. L.; Moir, R. W.; Wilcox, T. P.; Fink, J. H.</p> <p>1981-01-01</p> <p>A design for a 120-<span class="hlt">keV</span>, 2.3-MW, He-3 neutral beam injector for use on a D-(He-3) fusion reactor is described. The constraint that limits operating life when injecting He is its high sputtering rate. The sputtering is partly controlled by using an extra grid to prevent ion flow from the neutralizer duct to the electron suppressor grid, but a tradeoff between beam current and operating life is still required. Hollow grid wires functioning as mercury heat pipes cool the grid and enable steady state operation. Voltage holding and radiation effects on the acceleration grid structure are discussed. The vacuum system is also briefly described, and the use of a direct <span class="hlt">energy</span> converter to recapture <span class="hlt">energy</span> from unneutralized ions exiting the neutralizer is also analyzed. Of crucial importance to the technical feasibility of the (He-3)-burning reactor are the injector efficiency and cost; these are 53% and $5.5 million, respectively, when power supplies are included.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1981JFuE....1...69B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1981JFuE....1...69B"><span>Design study of a 120-<span class="hlt">keV</span>,3He neutral beam injector</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Blum, A. S.; Barr, W. L.; Dexter, W. L.; Fink, J. H.; Moir, R. W.; Wilcox, T. P.</p> <p>1981-01-01</p> <p>We describe a design for a 120-<span class="hlt">keV</span>, 2.3-MW,3He neutral beam injector for use on a D-3He fusion reactor. The constraint that limits operating life when injecting He is its high sputtering rate. The sputtering is partly controlled by using an extra grid to prevent ion flow from the neutralizer duct to the electron suppressor grid, but a tradeoff between beam current and operating life is still required. Hollow grid wires functioning as mercury heat pipes cool the grid and enable steady state operation. Voltage holding and radiation effects on the acceleration grid structure are discussed. We also briefly describe the vacuum system and analyze use of a direct <span class="hlt">energy</span> converter to recapture <span class="hlt">energy</span> from unneutralized ions exiting the neutralizer. Of crucial importance to the technical feasibility of the3He-burning reactor are the injector efficiency and cost; these are 53% and 5.5 million, respectively, when power supplies are included.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017EPJA...53...27P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EPJA...53...27P"><span>Study of the β- decay of 116m1In: A new interpretation of low-lying 0+ states in 116Sn</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Pore, J. L.; Cross, D. S.; Andreoiu, C.; Ashley, R.; Ball, G. C.; Bender, P. C.; Chester, A. S.; Diaz Varela, A.; Demand, G. A.; Dunlop, R.; Garnsworthy, A. B.; Garrett, P. E.; Hackman, G.; Hadinia, B.; Jigmeddorj, B.; Laffoley, A. T.; Liblong, A.; Kanungo, R.; Noakes, B.; Petrache, C. M.; Rajabali, M. M.; Starosta, K.; Svensson, C. E.; Voss, P. J.; Wang, Z. M.; Wood, J. L.; Yates, S. W.</p> <p>2017-02-01</p> <p>The 116Sn nucleus contains a collective rotational <span class="hlt">band</span> originating from proton π 2 p-2 h excitations across the proton Z=50 shell gap. Even though this nucleus has been extensively investigated in the past, there was still missing information on the low-<span class="hlt">energy</span> interband transitions connecting the intruder and normal structures. The low-lying structure of 116Sn was investigated through a high-statistics study of the β- decay of 116m1In with the 8π spectrometer and its ancillary detectors at TRIUMF. These measurements are critical in order to properly characterize the π 2 p-2 h rotational <span class="hlt">band</span>. Weak γ-decay branches are observed utilizing γ-γ coincidence spectroscopy methods, leading to the first direct observation of the 85 <span class="hlt">keV</span> 22+→ 03+ γ ray with a transition strength of B(E2) = 99.7(84) W.u. The analysis of these results strongly suggests that the 2027 <span class="hlt">keV</span> 03+ state should replace the previously assigned 1757 <span class="hlt">keV</span> 02+ state as the <span class="hlt">band</span>-head of the π 2 p-2 h rotational <span class="hlt">band</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 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('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/pages/biblio/1357545-fermi-observations-high-energy-gamma-ray-emission-from-grb','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1357545-fermi-observations-high-energy-gamma-ray-emission-from-grb"><span>FERMI observations of high-<span class="hlt">energy</span> gamma-ray emission from GRB 090217A</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Ackermann, M.; Ajello, M.; Baldini, L.; ...</p> <p>2010-06-22</p> <p>The Fermi observatory is advancing our knowledge of gamma-ray bursts (GRBs) through pioneering observations at high <span class="hlt">energies</span>, covering more than seven decades in <span class="hlt">energy</span> with the two on-board detectors, the Large Area Telescope (LAT) and the Gamma-ray Burst Monitor (GBM). Here, we report on the observation of the long GRB 090217A which triggered the GBM and has been detected by the LAT with a significance greater than 9σ. We present the GBM and LAT observations and on-ground analyses, including the time-resolved spectra and the study of the temporal profile from 8 <span class="hlt">keV</span> up to ~1 GeV. All spectra are wellmore » reproduced by a <span class="hlt">Band</span> model. We compare these observations to the first two LAT-detected, long bursts GRB 080825C and GRB 080916C. These bursts were found to have time-dependent spectra and exhibited a delayed onset of the high-<span class="hlt">energy</span> emission, which are not observed in the case of GRB 090217A. We discuss some theoretical implications for the high-<span class="hlt">energy</span> emission of GRBs.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018MNRAS.tmp.1434C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018MNRAS.tmp.1434C"><span>Searches for 3.5 <span class="hlt">keV</span> Absorption Features in Cluster AGN Spectra</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Conlon, Joseph P.</p> <p>2018-06-01</p> <p>We investigate possible evidence for a spectral dip around 3.5 <span class="hlt">keV</span> in central cluster AGNs, motivated by previous results for archival Chandra observations of the Perseus cluster and the general interest in novel spectral features around 3.5 <span class="hlt">keV</span> that may arise from dark matter physics. We use two deep Chandra observations of the Perseus and Virgo clusters that have recently been made public. In both cases, mild improvements in the fit (Δχ2 = 4.2 and Δχ2 = 2.5) are found by including such a dip at 3.5 <span class="hlt">keV</span> into the spectrum. A comparable result (Δχ2 = 6.5) is found re-analysing archival on-axis Chandra ACIS-S observations of the centre of the Perseus cluster.</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://hdl.handle.net/2060/20120014194','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20120014194"><span>When a Standard Candle Flickers: Crab Nebula Variations in Hard X-rays</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Wilson-Hodge, Collen A.; Cherry, M. L.; Case, G. L.; Baumgartner, W. H.; Beklen, E.; Bhat, P. N.; Briggs, M. S.; Camero-Arranz, A.; Chaplin, V.; Connaughton, V.; <a style="text-decoration: none; " href="javascript:void(0); " onClick="displayelement('author_20120014194'); toggleEditAbsImage('author_20120014194_show'); toggleEditAbsImage('author_20120014194_hide'); "> <img style="display:inline; width:12px; height:12px; " src="images/arrow-up.gif" width="12" height="12" border="0" alt="hide" id="author_20120014194_show"> <img style="width:12px; height:12px; display:none; " src="images/arrow-down.gif" width="12" height="12" border="0" alt="hide" id="author_20120014194_hide"></p> <p>2012-01-01</p> <p>RXTE played a crucial role in our surprising discovery that the Crab Nebula is variable in hard X-rays. In the first two years of science operations of the Fermi Gamma-ray Burst Monitor (GBM), August 2008-2010, a approx.7% (70 mcrab) decline was discovered in the overall Crab Nebula flux in the 15 - 50 <span class="hlt">keV</span> <span class="hlt">band</span>, measured with the Earth occultation technique. This decline was independently confirmed in the 15-50 <span class="hlt">keV</span> <span class="hlt">band</span> with four other instruments: the RXTE/PCA, Swift/BAT, INTEGRAL/IBIS, and INTEGRAL/SPI. The pulsed flux measured with RXTE/PCA from 1999-2010 was consistent with the pulsar spin-down, indicating that the observed changes were nebular. From 2001 to 2010, the Crab nebula flux measured with RXTE/PCA was particularly variable, changing by up to approx.3.5% per year in the 15-50 <span class="hlt">keV</span> <span class="hlt">band</span>. These variations were confirmed with INTEGRAL/SPI starting in 2003 and Swift/BAT starting in 2005. Before 2001 and since 2010, the Crab nebula 15-50 <span class="hlt">keV</span> flux measured with RXTE/PCA appeared more stable, varying by less than 2% per year. In this talk I will present Crab light curves including RXTE data for the entire 16-year mission in multiple <span class="hlt">energy</span> <span class="hlt">bands</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20040031731&hterms=gbm&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dgbm','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20040031731&hterms=gbm&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dgbm"><span>GLAST's GBM Burst Trigger</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Band, D.; Briggs, M.; Connaughton, V.; Kippen, M.; Preece, R.</p> <p>2003-01-01</p> <p>The GLAST Burst Monitor (GBM) will detect and localize bursts for the GLAST mission, and provide the spectral and temporal context in the traditional 10 <span class="hlt">keV</span> to 25 MeV <span class="hlt">band</span> for the high <span class="hlt">energy</span> observations by the Large Area Telescope (LAT). The GBM will use traditional rate triggers in up to three <span class="hlt">energy</span> <span class="hlt">bands</span>, and on a variety of timescales between 16 ms and 16 s.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1986PhRvA..34.1571K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1986PhRvA..34.1571K"><span>K-shell photoelectric cross sections for intermediate-Z elements at 26 <span class="hlt">keV</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kumar, Suresh; Singh, N.; Allawadhi, K. L.; Sood, B. S.</p> <p>1986-08-01</p> <p>Our earlier measurements of K-shell photoelectric cross sections for intermediate Z elements at 74 and 37 <span class="hlt">keV</span> have been extended to 26 <span class="hlt">keV</span> using external conversion x rays in Sn. The experimental results are found to show fairly good agreement with the theoretical values of Scofield.</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('http://adsabs.harvard.edu/abs/2005yCat..34280383D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2005yCat..34280383D"><span>VizieR Online Data Catalog: XMM-Newton Bright Serendipitous Survey (Della Ceca+, 2004)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Della Ceca, R.; Maccacaro, T.; Caccianiga, A.; Severgnini, P.; Braito, V.; Barcons, X.; Carrera, F. J.; Watson, M. G.; Tedds, J. A.; Brunner, H.; Lehmann, I.; Page, M. J.; Lamer, G.; Schwope, A.</p> <p>2005-09-01</p> <p>We present here "The XMM-Newton Bright Serendipitous Survey", composed of two flux-limited samples: the XMM-Newton Bright Source Sample (BSS, hereafter) and the XMM-Newton "Hard" Bright Source Sample (HBSS, hereafter) having a flux limit of fX~7x10-14erg/cm2/s in the 0.5-4.5<span class="hlt">keV</span> and 4.5-7.5<span class="hlt">keV</span> <span class="hlt">energy</span> <span class="hlt">band</span>, respectively. After discussing the main goals of this project and the survey strategy, we present the basic data on a complete sample of 400 X-ray sources (389 of them belong to the BSS, 67 to the HBSS with 56 X-ray sources in common) derived from the analysis of 237 suitable XMM-Newton fields (211 for the HBSS). At the flux limit of the survey we cover a survey area of 28.10 (25.17 for the HBSS) sq. deg. The extragalactic number-flux relationships (in the 0.5-4.5<span class="hlt">keV</span> and in the 4.5-7.5<span class="hlt">keV</span> <span class="hlt">energy</span> <span class="hlt">bands</span>) are in good agreement with previous and new results making us confident about the correctness of data selection and analysis. (5 data files).</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('http://adsabs.harvard.edu/abs/2013ApJ...769...22L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013ApJ...769...22L"><span>On the Spectral Hardening at gsim300 <span class="hlt">keV</span> in Solar Flares</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Li, G.; Kong, X.; Zank, G.; Chen, Y.</p> <p>2013-05-01</p> <p>It has long been noted that the spectra of observed continuum emissions in many solar flares are consistent with double power laws with a hardening at <span class="hlt">energies</span> gsim300 <span class="hlt">keV</span>. It is now widely believed that at least in electron-dominated events, the hardening in the photon spectrum reflects an intrinsic hardening in the source electron spectrum. In this paper, we point out that a power-law spectrum of electrons with a hardening at high <span class="hlt">energies</span> can be explained by the diffusive shock acceleration of electrons at a termination shock with a finite width. Our suggestion is based on an early analytical work by Drury et al., where the steady-state transport equation at a shock with a tanh profile was solved for a p-independent diffusion coefficient. Numerical simulations with a p-dependent diffusion coefficient show hardenings in the accelerated electron spectrum that are comparable with observations. One necessary condition for our proposed scenario to work is that high-<span class="hlt">energy</span> electrons resonate with the inertial range of the MHD turbulence and low-<span class="hlt">energy</span> electrons resonate with the dissipation range of the MHD turbulence at the acceleration site, and the spectrum of the dissipation range ~k -2.7. A ~k -2.7 dissipation range spectrum is consistent with recent solar wind observations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018NIMPA.877..220S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018NIMPA.877..220S"><span>Low <span class="hlt">energy</span> recoil detection with a spherical proportional counter</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Savvidis, I.; Katsioulas, I.; Eleftheriadis, C.; Giomataris, I.; Papaevangellou, T.</p> <p>2018-01-01</p> <p>We present results for the detection of low <span class="hlt">energy</span> nuclear recoils in the <span class="hlt">keV</span> <span class="hlt">energy</span> region, from measurements performed with the Spherical Proportional Counter (SPC). An 241Am-9Be fast neutron source is used in order to obtain neutron-nucleus elastic scattering events inside the gaseous volume of the detector. The detector performance in the <span class="hlt">keV</span> <span class="hlt">energy</span> region was measured by observing the 5.9 <span class="hlt">keV</span> line of a 55Fe X-ray source, with <span class="hlt">energy</span> resolution of 10% (σ). The toolkit GEANT4 was used to simulate the irradiation of the detector by an 241Am-9Be source, while SRIM was used to calculate the Ionization Quenching Factor (IQF), the simulation results are compared with the measurements. The potential of the SPC in low <span class="hlt">energy</span> recoil detection makes the detector a good candidate for a wide range of applications, including Supernova or reactor neutrino detection and Dark Matter (WIMP) searches (via coherent elastic scattering).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2001SPIE.4507..213P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2001SPIE.4507..213P"><span>Low-temperature high-Z gamma-detectors with very high <span class="hlt">energy</span> resolution</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Pobes, Carlos; Brofferio, Chiara; Bucci, Carlo; Cremonesi, Oliviero; Fiorini, Ettore; Giuliani, Andrea; Nucciotti, Angelo; Pavan, Maura; Pedretti, Marisa; Pessina, Gianluigi; Pirro, Stefano; Previtali, Ezio; Sisti, Monica; Vanzini, Marco; Zanotti, Luigi</p> <p>2001-12-01</p> <p>High-Z low-temperature calorimeters are developed by an Italian collaboration (Milano-Como-Gran Sasso Underground Laboratories) in order to search for rare nuclear events and Dark Matter massive candidates. They exhibit an excellent <span class="hlt">energy</span> resolution, close to that of Ge-diodes, but a much higher efficiency. Different high-Z materials were initially employed . A many-years optimisation work on tellurium oxide (TeO2) lead to impressive results: devices with total masses around 750 g present FWHM <span class="hlt">energy</span> resolutions on gamma-ray peaks ranging from 1 <span class="hlt">KeV</span> (close to the 5 <span class="hlt">KeV</span> <span class="hlt">energy</span> threshold) to 2.6 <span class="hlt">KeV</span> at 2615 <span class="hlt">KeV</span> (208Tl gamma line). A 3.2 <span class="hlt">KeV</span> FWHM <span class="hlt">energy</span> resolution was obtained at 5.4 MeV (210Po alpha line), which is by far the best one ever achieved with any alpha detector. These devices, operated at about 10 mK, consist of a TeO2 single crystal thermally coupled to a 50 mg Neutron Transmutation Doped (NTD) Ge crystal working as a temperature sensor. Special care was devoted to methods for response linearization and temporal stabilisation. Devices based on the same principle and specifically optimised could find applications in several fields like gamma-ray astrophysics, nuclear physics searches, environmental monitoring and radiation metrology.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2009AIPC.1126..351C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009AIPC.1126..351C"><span>Hard X-ray Flux from Low-Mass Stars in the Cygnus OB2 Association</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Caramazza, M.; Drake, J. J.; Micela, G.; Flaccomio, E.</p> <p>2009-05-01</p> <p>We investigate the X-ray emission in the 20-40 <span class="hlt">keV</span> <span class="hlt">band</span> expected from the flaring low-mass stellar population in Cygnus OB2 assuming that the observed soft X-ray emission is due to a superposition of flares and that the ratio of hard X-ray to soft X-ray emission is described by a scaling found for solar flares by Isola and co-workers. We estimate a low-mass stellar hard X-ray flux in the 20-40 <span class="hlt">keV</span> <span class="hlt">band</span> in the range ~7×1031-7×1033 erg/s and speculate the limit of this values. Hard X-ray emission could lie at a level not much below the current observed flux upper limits for Cygnus OB2. Simbol-X, with its broad <span class="hlt">energy</span> <span class="hlt">band</span> (10-100 <span class="hlt">keV</span>) and its sensitivity should be able to detect this emission and would provide insights into the hard X-ray production of flares on pre-main sequence stars.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017PhRvD..96k5019B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017PhRvD..96k5019B"><span>Common origin of 3.55 <span class="hlt">keV</span> x-ray line and gauge coupling unification with left-right dark matter</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Borah, Debasish; Dasgupta, Arnab; Patra, Sudhanwa</p> <p>2017-12-01</p> <p>We present a minimal left-right dark matter framework that can simultaneously explain the recently observed 3.55 <span class="hlt">keV</span> x-ray line from several galaxy clusters and gauge coupling unification at high <span class="hlt">energy</span> scale. Adopting a minimal dark matter strategy, we consider both left and right handed triplet fermionic dark matter candidates which are stable by virtue of a remnant Z2≃(-1 )B -L symmetry arising after the spontaneous symmetry breaking of left-right gauge symmetry to that of the standard model. A scalar bitriplet field is incorporated whose first role is to allow radiative decay of right handed triplet dark matter into the left handed one and a photon with <span class="hlt">energy</span> 3.55 <span class="hlt">keV</span>. The other role this bitriplet field at TeV scale plays is to assist in achieving gauge coupling unification at a high <span class="hlt">energy</span> scale within a nonsupersymmetric S O (10 ) model while keeping the scale of left-right gauge symmetry around the TeV corner. Apart from solving the neutrino mass problem and giving verifiable new contributions to neutrinoless double beta decay and charged lepton flavor violation, the model with TeV scale gauge bosons can also give rise to interesting collider signatures like diboson excess, dilepton plus two jets excess reported recently in the large hadron collider data.</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/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_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('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/24320422','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24320422"><span>A new concept of pencil beam dose calculation for 40-200 <span class="hlt">keV</span> photons using analytical dose kernels.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Bartzsch, Stefan; Oelfke, Uwe</p> <p>2013-11-01</p> <p>The advent of widespread kV-cone beam computer tomography in image guided radiation therapy and special therapeutic application of <span class="hlt">keV</span> photons, e.g., in microbeam radiation therapy (MRT) require accurate and fast dose calculations for photon beams with <span class="hlt">energies</span> between 40 and 200 <span class="hlt">keV</span>. Multiple photon scattering originating from Compton scattering and the strong dependence of the photoelectric cross section on the atomic number of the interacting tissue render these dose calculations by far more challenging than the ones established for corresponding MeV beams. That is why so far developed analytical models of kV photon dose calculations fail to provide the required accuracy and one has to rely on time consuming Monte Carlo simulation techniques. In this paper, the authors introduce a novel analytical approach for kV photon dose calculations with an accuracy that is almost comparable to the one of Monte Carlo simulations. First, analytical point dose and pencil beam kernels are derived for homogeneous media and compared to Monte Carlo simulations performed with the Geant4 toolkit. The dose contributions are systematically separated into contributions from the relevant orders of multiple photon scattering. Moreover, approximate scaling laws for the extension of the algorithm to inhomogeneous media are derived. The comparison of the analytically derived dose kernels in water showed an excellent agreement with the Monte Carlo method. Calculated values deviate less than 5% from Monte Carlo derived dose values, for doses above 1% of the maximum dose. The analytical structure of the kernels allows adaption to arbitrary materials and photon spectra in the given <span class="hlt">energy</span> range of 40-200 <span class="hlt">keV</span>. The presented analytical methods can be employed in a fast treatment planning system for MRT. In convolution based algorithms dose calculation times can be reduced to a few minutes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29787496','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29787496"><span>Quantitative Comparison of Virtual Monochromatic Images of Dual <span class="hlt">Energy</span> Computed Tomography Systems: Beam Hardening Artifact Correction and Variance in Computed Tomography Numbers: A Phantom Study.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wu, Rongli; Watanabe, Yoshiyuki; Satoh, Kazuhiko; Liao, Yen-Peng; Takahashi, Hiroto; Tanaka, Hisashi; Tomiyama, Noriyuki</p> <p>2018-05-21</p> <p>The aim of this study was to quantitatively compare the reduction in beam hardening artifact (BHA) and variance in computed tomography (CT) numbers of virtual monochromatic <span class="hlt">energy</span> (VME) images obtained with 3 dual-<span class="hlt">energy</span> computed tomography (DECT) systems at a given radiation dose. Five different iodine concentrations were scanned using dual-<span class="hlt">energy</span> and single-<span class="hlt">energy</span> (120 kVp) modes. The BHA and CT number variance were evaluated. For higher iodine concentrations, 40 and 80 mgI/mL, BHA on VME imaging was significantly decreased when the <span class="hlt">energy</span> was higher than 50 <span class="hlt">keV</span> (P = 0.003) and 60 <span class="hlt">keV</span> (P < 0.001) for GE, higher than 80 <span class="hlt">keV</span> (P < 0.001) and 70 <span class="hlt">keV</span> (P = 0.002) for Siemens, and higher than 40 <span class="hlt">keV</span> (P < 0.001) and 60 <span class="hlt">keV</span> (P < 0.001) for Toshiba, compared with single-<span class="hlt">energy</span> CT imaging. Virtual monochromatic <span class="hlt">energy</span> imaging can decrease BHA and improve CT number accuracy in different dual-<span class="hlt">energy</span> computed tomography systems, depending on <span class="hlt">energy</span> levels and iodine concentrations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1436659-ultralow-energy-calibration-lux-detector-using-xe127-electron-capture','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1436659-ultralow-energy-calibration-lux-detector-using-xe127-electron-capture"><span>Ultralow <span class="hlt">energy</span> calibration of LUX detector using Xe 127 electron capture</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Akerib, D. S.; Alsum, S.; Araújo, H. M.; ...</p> <p>2017-12-01</p> <p>We report an absolute calibration of the ionization yields(more » $$\\textit{Q$$_y$})$ and fluctuations for electronic recoil events in liquid xenon at discrete <span class="hlt">energies</span> between 186 eV and 33.2 <span class="hlt">keV</span>. The average electric field applied across the liquid xenon target is 180 V/cm. The data are obtained using low <span class="hlt">energy</span> $$^{127}$$Xe electron capture decay events from the 95.0-day first run from LUX (WS2013) in search of Weakly Interacting Massive Particles (WIMPs). The sequence of gamma-ray and X-ray cascades associated with $$^{127}$$I de-excitations produces clearly identified 2-vertex events in the LUX detector. We observe the K- (binding <span class="hlt">energy</span>, 33.2 <span class="hlt">keV</span>), L- (5.2 <span class="hlt">keV</span>), M- (1.1 <span class="hlt">keV</span>), and N- (186 eV) shell cascade events and verify that the relative ratio of observed events for each shell agrees with calculations. The N-shell cascade analysis includes single extracted electron (SE) events and represents the lowest-<span class="hlt">energy</span> electronic recoil $$\\textit{in situ}$$ measurements that have been explored in liquid xenon.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017ApPhL.111u3502K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017ApPhL.111u3502K"><span>A tapered multi-gap multi-aperture pseudospark-sourced electron gun based X-<span class="hlt">band</span> slow wave oscillator</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kumar, N.; Lamba, R. P.; Hossain, A. M.; Pal, U. N.; Phelps, A. D. R.; Prakash, R.</p> <p>2017-11-01</p> <p>The experimental study of a tapered, multi-gap, multi-aperture pseudospark-sourced electron gun based X-<span class="hlt">band</span> plasma assisted slow wave oscillator is presented. The designed electron gun is based on the pseudospark discharge concept and has been used to generate a high current density and high <span class="hlt">energy</span> electron beam simultaneously. The distribution of apertures has been arranged such that the field penetration potency inside the backspace of the hollow-cathode is different while passing through the tapered gap region. This leads to non-concurrent ignition of the discharge through all the channels which is, in general, quite challenging in the case of multi-aperture plasma cathode electron gun geometries. Multiple and successive hollow cathode phases are reported from this electron gun geometry, which have been confirmed using simulations. This geometry also has led to the achievement of ˜71% fill factor inside the slow wave oscillator for an electron beam of <span class="hlt">energy</span> of 20 <span class="hlt">keV</span> and a beam current density in the range of 115-190 A/cm2 at a working argon gas pressure of 18 Pa. The oscillator has generated broadband microwave output in the frequency range of 10-11.7 GHz with a peak power of ˜10 kW for ˜50 ns.</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. Additionally, 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('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('https://www.osti.gov/biblio/21325808-experimental-results-modeling-low-heat-capacity-tes-microcalorimeters-soft-ray-spectroscopy','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/21325808-experimental-results-modeling-low-heat-capacity-tes-microcalorimeters-soft-ray-spectroscopy"><span>Experimental Results and Modeling of Low-Heat-Capacity TES Microcalorimeters for Soft-X-ray 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>Eckart, Megan E.; NASA Postdoctoral Program Fellow; Adams, Joseph S.</p> <p>2009-12-16</p> <p>Transition-edge-sensor (TES) X-ray microcalorimeters have mostly been targeted at mid-<span class="hlt">band</span> <span class="hlt">energies</span> from 0.05-10 <span class="hlt">keV</span> and high <span class="hlt">energies</span> to above 100 <span class="hlt">keV</span>. However, many other optimizations are possible. Here we present results from devices optimized for soft X-ray applications. For spectroscopy below 1 <span class="hlt">keV</span>, the X-ray stopping power and heat capacity (C) of the TES itself are high enough that we can omit a separate absorber. The resulting devices have low C and the best-achievable <span class="hlt">energy</span> resolution should be under 1 eV. We are interested in pursuing such devices primarily for astrophysical applications and laboratory astrophysics at LLNL's Electron-Beam Ion Trap.more » To this end, we have studied arrays in which 'bare' TESs are interspersed with broad-<span class="hlt">band</span> pixels that have absorbers. By extending the absorbers to cover the area where the leads contact the low-<span class="hlt">energy</span> pixels, we have eliminated a significant source of non-Gaussian detector response. The bare devices are in a different regime from our typical devices in that C is ten times lower and the conductance to the bath is four times lower. We have explored this regime through simultaneous fitting of noise and impedance data. These data cannot be fit by the simple model we employ to describe our typical broad-<span class="hlt">band</span> devices. In this contribution we present X-ray spectra and the results from modeling.« less</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.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('http://adsabs.harvard.edu/abs/2018PhRvD..97i2007A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PhRvD..97i2007A"><span>Signal yields of <span class="hlt">keV</span> electronic recoils and their discrimination from nuclear recoils in liquid xenon</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Aprile, E.; Aalbers, J.; Agostini, F.; Alfonsi, M.; Amaro, F. D.; Anthony, M.; Arneodo, F.; Barrow, P.; Baudis, L.; Bauermeister, B.; Benabderrahmane, M. L.; Berger, T.; Breur, P. A.; Brown, A.; Brown, E.; Bruenner, S.; Bruno, G.; Budnik, R.; Bütikofer, L.; Calvén, J.; Cardoso, J. M. R.; Cervantes, M.; Cichon, D.; Coderre, D.; Colijn, A. P.; Conrad, J.; Cussonneau, J. P.; Decowski, M. P.; de Perio, P.; di Gangi, P.; di Giovanni, A.; Diglio, S.; Eurin, G.; Fei, J.; Ferella, A. D.; Fieguth, A.; Fulgione, W.; Gallo Rosso, A.; Galloway, M.; Gao, F.; Garbini, M.; Geis, C.; Goetzke, L. W.; Grandi, L.; Greene, Z.; Grignon, C.; Hasterok, C.; Hogenbirk, E.; Howlett, J.; Itay, R.; Kaminsky, B.; Kazama, S.; Kessler, G.; Kish, A.; Landsman, H.; Lang, R. F.; Lellouch, D.; Levinson, L.; Lin, Q.; Lindemann, S.; Lindner, M.; Lombardi, F.; Lopes, J. A. M.; Mahlstedt, J.; Manfredini, A.; Maris, I.; Marrodán Undagoitia, T.; Masbou, J.; Massoli, F. V.; Masson, D.; Mayani, D.; Messina, M.; Micheneau, K.; Molinario, A.; Morâ, K.; Murra, M.; Naganoma, J.; Ni, K.; Oberlack, U.; Pakarha, P.; Pelssers, B.; Persiani, R.; Piastra, F.; Pienaar, J.; Pizzella, V.; Piro, M.-C.; Plante, G.; Priel, N.; Ramírez García, D.; Rauch, L.; Reichard, S.; Reuter, C.; Rizzo, A.; Rupp, N.; Saldanha, R.; Dos Santos, J. M. F.; Sartorelli, G.; Scheibelhut, M.; Schindler, S.; Schreiner, J.; Schumann, M.; Scotto Lavina, L.; Selvi, M.; Shagin, P.; Shockley, E.; Silva, M.; Simgen, H.; Sivers, M. V.; Stein, A.; Thers, D.; Tiseni, A.; Trinchero, G.; Tunnell, C.; Vargas, M.; Wang, H.; Wang, Z.; Wei, Y.; Weinheimer, C.; Wittweg, C.; Wulf, J.; Ye, J.; Zhang, Y.; Zhu, T.; Xenon Collaboration</p> <p>2018-05-01</p> <p>We report on the response of liquid xenon to low <span class="hlt">energy</span> electronic recoils below 15 <span class="hlt">keV</span> from beta decays of tritium at drift fields of 92 V /cm , 154 V /cm and 366 V /cm using the XENON100 detector. A data-to-simulation fitting method based on Markov Chain Monte Carlo is used to extract the photon yields and recombination fluctuations from the experimental data. The photon yields measured at the two lower fields are in agreement with those from literature; additional measurements at a higher field of 366 V /cm are presented. The electronic and nuclear recoil discrimination as well as its dependence on the drift field and photon detection efficiency are investigated at these low <span class="hlt">energies</span>. The results provide new measurements in the <span class="hlt">energy</span> region of interest for dark matter searches using liquid xenon.</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('https://www.ncbi.nlm.nih.gov/pubmed/12382856','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/12382856"><span>Thermoluminescent response of LiF:Mg,Ti to 20 <span class="hlt">keV</span> electrons.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Mercado-Uribe, H; Brandan, M E</p> <p>2002-01-01</p> <p>The thermoluminescence response of LiF:Mg,Ti (TLD-100) to 20 <span class="hlt">keV</span> electrons from a scanning electron microscope has been measured. Radiochromic dye films previously calibrated were used to determine the fluence incident on TLD-100 chips. The procedure for irradiation and glow curve deconvolution was adhered to the protocols previously determined in our laboratory for gamma rays and heavy charged particles. The response at electron fluences higher than 4 x 10(10) cm(-2) is supralinear, due to the increasingly relevant contribution of the high temperature peaks. The relative contribution of the high temperature peaks to the TL signal is abnormally small, about half that observed in gamma irradiation and four times smaller than what has been measured in low-<span class="hlt">energy</span> X ray exposure.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017Icar..288..284G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017Icar..288..284G"><span>The Mars diffuse aurora: A model of ultraviolet and visible emissions</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gérard, J.-C.; Soret, L.; Shematovich, V. I.; Bisikalo, D. V.; Bougher, S. W.</p> <p>2017-05-01</p> <p>A new type of Martian aurora, characterized by an extended spatial distribution, an altitude lower than the discrete aurora and electron precipitation up to 200 <span class="hlt">keV</span> has been observed following solar activity on several occasions from the MAVEN spacecraft. We describe the results of Monte Carlo simulations of the production of several ultraviolet and violet auroral emissions for initial electron <span class="hlt">energies</span> extending from 0.25 to 200 <span class="hlt">keV</span>. These include the CO2+ ultraviolet doublet (UVD) at 288.3 and 289.6 nm and the Fox-Duffendack-Barker (FDB) <span class="hlt">bands</span>, CO Cameron and Fourth Positive <span class="hlt">bands</span>, OI 130.4 and 297.2 nm and CI 156.1 nm and 165.7 nm multiplets. We calculate the nadir and limb production rates of several of these emissions for a unit precipitated <span class="hlt">energy</span> flux. Our results indicate that electrons in the range 50-200 <span class="hlt">keV</span> produce maximum CO2+ UVD emission below 75 km, in agreement with the MAVEN observations. We calculate the efficiency of photon production per unit precipitated electron power. The strongest emissions are the CO2+ FDB, UVD and CO Cameron <span class="hlt">bands</span> and the oxygen emission at 297.2 nm. The metastable a 3Π state which radiates the Cameron <span class="hlt">bands</span> is deactivated by collisions below about 110 km. As a consequence, we show that the Cameron <span class="hlt">band</span> emission is expected to peak at a higher altitude than the CO2+ UVD and FDB <span class="hlt">bands</span>. Collisional quenching also causes the intensity ratio of the CO2+ UVD to CO Cameron <span class="hlt">bands</span> to increase below ∼100 km in the energetic diffuse aurora.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1343044-nustar-observations-magnetar','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1343044-nustar-observations-magnetar"><span>NuSTAR OBSERVATIONS OF MAGNETAR 1E 1048.1–5937</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Yang, C.; Archibald, R. F.; Vogel, J. K.; ...</p> <p>2016-10-28</p> <p>We report on simultaneous Nuclear Spectroscopic Telescope Array (NuSTAR) and XMM-Newton observations of the magnetar 1E 1048.1-5937, along with Rossi X-ray Timing Explorer (RXTE) data for the same source. The NuSTAR data provide a clear detection of this magnetar’s persistent emission up to 20 <span class="hlt">keV</span>. We detect a previously unreported small secondary peak in the average pulse profile in the 7–10 <span class="hlt">keV</span> <span class="hlt">band</span>, which grows to an amplitude comparable to that of the main peak in the 10–20 <span class="hlt">keV</span> <span class="hlt">band</span>. We show using RXTE data that this secondary peak is likely transient. We find that the pulsed fraction increases withmore » <span class="hlt">energy</span> from a value of ~0.55 at ~2 <span class="hlt">keV</span> to a value of ~0.75 near 8 <span class="hlt">keV</span> but shows evidence of decreasing at higher <span class="hlt">energies</span>. After filtering out multiple bright X-ray bursts during the observation, we find that the phase-averaged spectrum from combined NuSTAR and XMM data is well described by an absorbed double blackbody plus power-law model, with no evidence for the spectral turn-up near ~10 <span class="hlt">keV</span> as has been seen in some other magnetars. Our data allow us to rule out a spectral turn-up similar to those seen in magnetars 4U 0142+61 and 1E 2259+586 of ΔΓ≳2, where ΔΓ is the difference between the softband and hard-<span class="hlt">band</span> photon indexes. The lack of spectral turn-up is consistent with what has been observed from an active subset of magnetars given previously reported trends suggesting that the degree of spectral turn-up is correlated with spin-down rate and/or spin-inferred magnetic field.« less</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('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://hdl.handle.net/2060/19770027140','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19770027140"><span>Charge states of low <span class="hlt">energy</span> ions from the sun. Ph.D. Thesis</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Sciambi, R. K.</p> <p>1975-01-01</p> <p>Measurements of ionization states and <span class="hlt">energy</span> spectra of carbon, oxygen, and iron accelerated in ten solar flare particle events are reported, for <span class="hlt">energies</span> between 15 <span class="hlt">keV</span> per nucleon and 600 <span class="hlt">keV</span> per nucleon. The ionization states were remarkably constant from flare to flare, despite great variations in other event parameters. The mean ionization state for carbon was 5.7, for oxygen 6.2, and for iron 11.7, values which are similar to the respective ionization states in the solar wind. The time profile of the He/C+N+O ratio was examined, and it was found that the ratio was small early in the event, and increased with time. The <span class="hlt">energy</span> spectra of the medium ions showed a flattening below 100 <span class="hlt">keV</span> per nucleon, which was highly correlated with event size as measured by the event averaged flux of 130 to 220 <span class="hlt">keV</span> protons.</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>