Sample records for na osnove cai2

  1. Structural basis of Na(+)-independent and cooperative substrate/product antiport in CaiT.

    PubMed

    Schulze, Sabrina; Köster, Stefan; Geldmacher, Ulrike; Terwisscha van Scheltinga, Anke C; Kühlbrandt, Werner

    2010-09-09

    Transport of solutes across biological membranes is performed by specialized secondary transport proteins in the lipid bilayer, and is essential for life. Here we report the structures of the sodium-independent carnitine/butyrobetaine antiporter CaiT from Proteus mirabilis (PmCaiT) at 2.3-A and from Escherichia coli (EcCaiT) at 3.5-A resolution. CaiT belongs to the family of betaine/carnitine/choline transporters (BCCT), which are mostly Na(+) or H(+) dependent, whereas EcCaiT is Na(+) and H(+) independent. The three-dimensional architecture of CaiT resembles that of the Na(+)-dependent transporters LeuT and BetP, but in CaiT a methionine sulphur takes the place of the Na(+) ion to coordinate the substrate in the central transport site, accounting for Na(+)-independent transport. Both CaiT structures show the fully open, inward-facing conformation, and thus complete the set of functional states that describe the alternating access mechanism. EcCaiT contains two bound butyrobetaine substrate molecules, one in the central transport site, the other in an extracellular binding pocket. In the structure of PmCaiT, a tryptophan side chain occupies the transport site, and access to the extracellular site is blocked. Binding of both substrates to CaiT reconstituted into proteoliposomes is cooperative, with Hill coefficients up to 1.7, indicating that the extracellular site is regulatory. We propose a mechanism whereby the occupied regulatory site increases the binding affinity of the transport site and initiates substrate translocation.

  2. Na/Ca exchange in the basolateral membrane of the A6 cell monolayer: role in Cai homeostasis.

    PubMed

    Brochiero, E; Raschi, C; Ehrenfeld, J

    1995-05-01

    The presence of a Na/Ca exchanger in A6 cells was investigated by measuring intracellular calcium (Cai) fluctuations and the 45Ca fluxes through the basolateral membranes (blm) of the cell monolayer. Removal of Na+ from the medium produced a transient increase in Cai followed by a regulatory phase returning Cai to control levels in 3-4 min, this phase being greatly accelerated (< 60 s) by NaCl addition (apparent Km of approximately 5 mM Na+). The Cai increase was only found with the Na(+)-free medium on the basolateral side of the cell monolayer. A twofold increase in the 45Ca influx was observed under these conditions. In Ca(2+)- depleted cells, the initial Cai increase after Ca2+ addition to the medium was greater when the putative Na/Ca exchanger was not functioning (i.e. in a Na(+)-free medium). 45Ca effluxes through the blm of the monolayer were greatly and transiently increased by a Na(+)-free medium on the serosal side and blocked by orthovanadate (1 mM). The Cai increased induced by a hypo-osmotic shock was greater in cells bathed in a Na(+)-medium, conditions expected to block the activity of the Na/Ca exchanger. These findings support the hypothesis that a Na/Ca exchanger is present on the blm of A6 cells and affirm its role in Cai homeostasis in steady-state conditions and following osmotic shock. In addition, a Ca2+ pump also located on the blm and Ca2+ stores sensitive to inositol 1,4,5-trisphosphate were found to be implicated in Cai homeostasis.

  3. The impact of the use of different satellite data as training data against GOSAT-2 CAI-2 L2 cloud discrimination

    NASA Astrophysics Data System (ADS)

    Oishi, Y.; Ishida, H.; Nakajima, T. Y.

    2016-12-01

    Greenhouse gases Observing SATellite-2 (GOSAT-2) will be launched in fiscal 2017 to determine atmospheric concentrations of greenhouse gases, such as CO2, CH4, and CO. GOSAT-2 will be equipped with two sensors: the Thermal and Near-infrared Sensor for Carbon Observation (TANSO)-Fourier Transform Spectrometer-2 (FTS-2) and TANSO-Cloud and Aerosol Imager-2 (CAI-2). CAI-2 is a push-broom imaging sensor that has forward- and backward-looking bands for observing the optical properties of aerosols and clouds, and for monitoring the status of urban air pollution and transboundary air pollution over oceans. An important role of CAI-2 is to perform cloud discrimination in each direction. The Cloud and Aerosol Unbiased Decision Intellectual Algorithm (CLAUDIA1), which applies sequential threshold tests to features, has been used in GOSAT CAI L2 cloud flag processing. If CLAUDIA1 used with CAI-2, it is necessary to optimize the thresholds in accordance with CAI-2. Meanwhile, CLAUDIA3 using support vector machines (SVM), which is a supervised pattern recognition method, was developed for GOSAT-2 CAI-2 L2 cloud discrimination processing. Thus, CLAUDIA3 can automatically find the optimized boundary between clear and cloudy. Improvement of the CLAUDIA3 used with CAI (CLAUDIA3-CAI) has carried out and is still continuing. In this study we compared results of CLAUDIA3-CAI using Terra MODIS data and GOSAT CAI data as training data to clarify the impact of the use of different satellite data as training data against GOSAT-2 CAI-2 L2 cloud discrimination. We will present our latest results.

  4. A ryanodine receptor-dependent Ca(i)(2+) asymmetry at Hensen's node mediates avian lateral identity.

    PubMed

    Garic-Stankovic, Ana; Hernandez, Marcos; Flentke, George R; Zile, Maija H; Smith, Susan M

    2008-10-01

    In mouse, the establishment of left-right (LR) asymmetry requires intracellular calcium (Ca(i)(2+)) enrichment on the left of the node. The use of Ca(i)(2+) asymmetry by other vertebrates, and its origins and relationship to other laterality effectors are largely unknown. Additionally, the architecture of Hensen's node raises doubts as to whether Ca(i)(2+) asymmetry is a broadly conserved mechanism to achieve laterality. We report here that the avian embryo uses a left-side enriched Ca(i)(2+) asymmetry across Hensen's node to govern its lateral identity. Elevated Ca(i)(2+) was first detected along the anterior node at early HH4, and its emergence and left-side enrichment by HH5 required both ryanodine receptor (RyR) activity and extracellular calcium, implicating calcium-induced calcium release (CICR) as the novel source of the Ca(i)(2+). Targeted manipulation of node Ca(i)(2+) randomized heart laterality and affected nodal expression. Bifurcation of the Ca(i)(2+) field by the emerging prechordal plate may permit the independent regulation of LR Ca(i)(2+) levels. To the left of the node, RyR/CICR and H(+)V-ATPase activity sustained elevated Ca(i)(2+). On the right, Ca(i)(2+) levels were actively repressed through the activities of H(+)K(+) ATPase and serotonin-dependent signaling, thus identifying a novel mechanism for the known effects of serotonin on laterality. Vitamin A-deficient quail have a high incidence of situs inversus hearts and had a reversed calcium asymmetry. Thus, Ca(i)(2+) asymmetry across the node represents a more broadly conserved mechanism for laterality among amniotes than had been previously believed.

  5. CAI Update: So You Want to Do CAI?

    ERIC Educational Resources Information Center

    Bagley, Carole

    1979-01-01

    Provides necessary characteristics to consider when selecting a CAI system plus a list of costs and capabilities available with the better known CAI systems. Characteristics of major CAI systems are presented in three categories--large/maxi, mini, and micro systems--in chart form. (JEG)

  6. Nanocomposite Phosphor Consisting of CaI2:Eu2+ Single Nanocrystals Embedded in Crystalline SiO2.

    PubMed

    Daicho, Hisayoshi; Iwasaki, Takeshi; Shinomiya, Yu; Nakano, Akitoshi; Sawa, Hiroshi; Yamada, Wataru; Matsuishi, Satoru; Hosono, Hideo

    2017-11-29

    High luminescence efficiency is obtained in halide- and chalcogenide-based phosphors, but they are impractical because of their poor chemical durability. Here we report a halide-based nanocomposite phosphor with excellent luminescence efficiency and sufficient durability for practical use. Our approach was to disperse luminescent single nanocrystals of CaI 2 :Eu 2+ in a chemically stable, translucent crystalline SiO 2 matrix. Using this approach, we successfully prepared a nanocomposite phosphor by means of self-organization through a simple solid-state reaction. Single nanocrystals of 6H polytype (thr notation) CaI 2 :Eu 2+ with diameters of about 50 nm could be generated not only in a SiO 2 amorphous powder but also in a SiO 2 glass plate. The nanocomposite phosphor formed upon solidification of molten CaI 2 left behind in the crystalline SiO 2 that formed from the amorphous SiO 2 under the influence of a CaI 2 flux effect. The resulting nanocomposite phosphor emitted brilliant blue luminescence with an internal quantum efficiency up to 98% upon 407 nm violet excitation. We used cathodoluminescence microscopy, scanning transmission electron microscopy, and Rietveld refinement of the X-ray diffraction patterns to confirm that the blue luminescence was generated only by the CaI 2 :Eu 2+ single nanocrystals. The phosphor was chemically durable because the luminescence sites were embedded in the crystalline SiO 2 matrix. The phosphor is suitable for use in near-ultraviolet light-emitting diodes. The concept for this nanocomposite phosphor can be expected to be effective for improvements in the practicality of poorly durable materials such as halides and chalcogenides.

  7. Preliminary verification for application of a support vector machine-based cloud detection method to GOSAT-2 CAI-2

    NASA Astrophysics Data System (ADS)

    Oishi, Yu; Ishida, Haruma; Nakajima, Takashi Y.; Nakamura, Ryosuke; Matsunaga, Tsuneo

    2018-05-01

    The Greenhouse Gases Observing Satellite (GOSAT) was launched in 2009 to measure global atmospheric CO2 and CH4 concentrations. GOSAT is equipped with two sensors: the Thermal And Near infrared Sensor for carbon Observations (TANSO)-Fourier transform spectrometer (FTS) and TANSO-Cloud and Aerosol Imager (CAI). The presence of clouds in the instantaneous field of view of the FTS leads to incorrect estimates of the concentrations. Thus, the FTS data suspected to have cloud contamination must be identified by a CAI cloud discrimination algorithm and rejected. Conversely, overestimating clouds reduces the amount of FTS data that can be used to estimate greenhouse gas concentrations. This is a serious problem in tropical rainforest regions, such as the Amazon, where the amount of useable FTS data is small because of cloud cover. Preparations are continuing for the launch of the GOSAT-2 in fiscal year 2018. To improve the accuracy of the estimates of greenhouse gases concentrations, we need to refine the existing CAI cloud discrimination algorithm: Cloud and Aerosol Unbiased Decision Intellectual Algorithm (CLAUDIA1). A new cloud discrimination algorithm using a support vector machine (CLAUDIA3) was developed and presented in another paper. Although the use of visual inspection of clouds as a standard for judging is not practical for screening a full satellite data set, it has the advantage of allowing for locally optimized thresholds, while CLAUDIA1 and -3 use common global thresholds. Thus, the accuracy of visual inspection is better than that of these algorithms in most regions, with the exception of snow- and ice-covered surfaces, where there is not enough spectral contrast to identify cloud. In other words, visual inspection results can be used as truth data for accuracy evaluation of CLAUDIA1 and -3. For this reason visual inspection can be used for the truth metric for the cloud discrimination verification exercise. In this study, we compared CLAUDIA1-CAI and

  8. Astrophysics of CAI formation as revealed by silicon isotope LA-MC-ICPMS of an igneous CAI

    NASA Astrophysics Data System (ADS)

    Shahar, Anat; Young, Edward D.

    2007-05-01

    Silicon isotope ratios of a typical CAI from the Leoville carbonaceous chondrite, obtained in situ by laser ablation MC-ICPMS, together with existing 25Mg/ 24Mg data, reveal a detailed picture of the astrophysical setting of CAI melting and subsequent heating. Models for the chemical and isotopic effects of evaporation of the molten CAI are used to produce a univariant relationship between PH 2 and time during melting. The result shows that this CAI was molten for a cumulative time of no more than 70 days and probably less than 15 days depending on temperature. The object could have been molten for an integrated time of just a few hours if isotope ratio zoning was eliminated after melting by high subsolidus temperatures (e.g., > 1300 K) for ˜ 500 yr. In all cases subsolidus heating sufficient to produce diffusion-limited isotope fractionation at the margin of the solidified CAI is required. These stable isotope data point to a two-stage history for this igneous CAI involving melting for a cumulative timescale of hours to months followed by subsolidus heating for years to hundreds of years. The thermobarometric history deduced from combining Si and Mg isotope ratio data implicates thermal processing in the disk, perhaps by passage through shockwaves, following melting. This study underscores the direct link between the meaning of stable isotope ratio zoning, or lack thereof, and the inferred astrophysical setting of melting and subsequent processing of CAIs.

  9. CAI and Developmental Education.

    ERIC Educational Resources Information Center

    Anderson, Rick

    This paper discusses the problems and achievements of computer assisted instruction (CAI) projects at University College, University of Cincinnati. The most intensive use of CAI on campus, the CAI Lab, is part of the Developmental Education Center's effort to serve students who lack mastery of basic college-level skills in mathematics and English.…

  10. Copyright and CAI.

    ERIC Educational Resources Information Center

    Kearsley, G.P.; Hunka, S.

    The application of copyright laws to Computer Assisted Instruction (CAI) is not a simple matter of extending traditional literary practices because of the legal complications introduced by the use of computers to store and reproduce materials. In addition, CAI courseware poses some new problems for the definitions of educational usage. Some…

  11. Ca-Fe and Alkali-Halide Alteration of an Allende Type B CAI: Aqueous Alteration in Nebular or Asteroidal Settings

    NASA Technical Reports Server (NTRS)

    Ross, D. K.; Simon, J. I.; Simon, S. B.; Grossman, L.

    2012-01-01

    Ca-Fe and alkali-halide alteration of CAIs is often attributed to aqueous alteration by fluids circulating on asteroidal parent bodies after the various chondritic components have been assembled, although debate continues about the roles of asteroidal vs. nebular modification processes [1-7]. Here we report de-tailed observations of alteration products in a large Type B2 CAI, TS4 from Allende, one of the oxidized subgroup of CV3s, and propose a speculative model for aqueous alteration of CAIs in a nebular setting. Ca-Fe alteration in this CAI consists predominantly of end-member hedenbergite, end-member andradite, and compositionally variable, magnesian high-Ca pyroxene. These phases are strongly concentrated in an unusual "nodule" enclosed within the interior of the CAI (Fig. 1). The Ca, Fe-rich nodule superficially resembles a clast that pre-dated and was engulfed by the CAI, but closer inspection shows that relic spinel grains are enclosed in the nodule, and corroded CAI primary phases interfinger with the Fe-rich phases at the nodule s margins. This CAI also contains abundant sodalite and nepheline (alkali-halide) alteration that occurs around the rims of the CAI, but also penetrates more deeply into the CAI. The two types of alteration (Ca-Fe and alkali-halide) are adjacent, and very fine-grained Fe-rich phases are associated with sodalite-rich regions. Both types of alteration appear to be replacive; if that is true, it would require substantial introduction of Fe, and transport of elements (Ti, Al and Mg) out of the nodule, and introduction of Na and Cl into alkali-halide rich zones. Parts of the CAI have been extensively metasomatized.

  12. Experience with the CAIS

    NASA Technical Reports Server (NTRS)

    Tighe, Michael F.

    1986-01-01

    Intermetrics' experience is that the Ada package construct, which allows separation of specification and implementation allows specification of a CAIS that is transportable across varying hardware and software bases. Additionally, the CAIS is an excellent basis for providing operating system functionality to Ada applications. By allowing the Byron APSE to be moved easily from system to system, and allowing significant re-writes of underlying code. Ada and the CAIS provide portability as well as transparency to change at the application operating system interface level.

  13. CAI System Costs: Present and Future.

    ERIC Educational Resources Information Center

    Pressman, Israel; Rosenbloom, Bruce

    1984-01-01

    Discusses costs related to providing computer assisted instruction (CAI), considering hardware, software, user training, maintenance, and installation. Provides an example of the total cost of CAI broken down into these categories, giving an adjusted yearly cost. Projects future trends and costs of CAI as well as cost savings possibilities. (JM)

  14. O, Mg, and Si isotope distributions in the complex ultrarefractory CAI Efremovka 101.1: Assimilation of ultrarefractory, FUN, and regular CAI precursors

    NASA Astrophysics Data System (ADS)

    Aléon, Jérôme; Marin-Carbonne, Johanna; McKeegan, Kevin D.; El Goresy, Ahmed

    2018-07-01

    Oxygen, magnesium, and silicon isotopic compositions in the mineralogically complex, ultrarefractory (UR) calcium-aluminum-rich inclusion (CAI) E101.1 from the reduced CV3 chondrite Efremovka confirm that E101.1 is a compound CAI composed of several lithological units that were once individual CAIs, free-floating in the solar protoplanetary disk. Each precursor unit was found to have had its own thermal history prior to being captured and incorporated into the partially molten host CAI. Four major lithological units can be distinguished on the basis of their isotopic compositions. (1) Al-diopside-rich sinuous fragments, hereafter sinuous pyroxene, are 16O-rich (Δ17O ≤ -20‰) and have light Mg and Si isotopic compositions with mass fractionation down to -3.5‰/amu for both isotopic systems. We attribute these peculiar isotopic compositions to kinetic effects during condensation out of thermal equilibrium. (2) Spinel clusters are 16O-rich (Δ17O ∼ -22‰) and have Mg isotope systematics consistent with extensive equilibration with the host melt. This includes (i) δ25Mg values varying between + 2.6‰ and + 6.5‰ close to the typical value of host melilite at ∼+5‰, and (ii) evidence for exchange of radiogenic 26Mg with adjacent melilite as indicated by Al/Mg systematics. The spinel clusters may represent fine-grained spinel-rich proto-CAIs captured, partially melted, and recrystallized in the host melt. Al/Mg systematics indicate that both the sinuous pyroxene fragments and spinel clusters probably had canonical or near-canonical 26Al contents before partial equilibration. (3) The main CAI host (Δ17O ≤ -2‰) had a complex thermal history partially obscured by subsequent capture and assimilation events. Its formation, referred to as the "cryptic" stage, could have resulted from the partial melting and crystallization of a 16O-rich precursor that underwent 16O-depletion and a massive evaporation event characteristic of F and FUN CAIs (Fractionated with

  15. Numerical simulation and validation of SI-CAI hybrid combustion in a CAI/HCCI gasoline engine

    NASA Astrophysics Data System (ADS)

    Wang, Xinyan; Xie, Hui; Xie, Liyan; Zhang, Lianfang; Li, Le; Chen, Tao; Zhao, Hua

    2013-02-01

    SI-CAI hybrid combustion, also known as spark-assisted compression ignition (SACI), is a promising concept to extend the operating range of CAI (Controlled Auto-Ignition) and achieve the smooth transition between spark ignition (SI) and CAI in the gasoline engine. In this study, a SI-CAI hybrid combustion model (HCM) has been constructed on the basis of the 3-Zones Extended Coherent Flame Model (ECFM3Z). An ignition model is included to initiate the ECFM3Z calculation and induce the flame propagation. In order to precisely depict the subsequent auto-ignition process of the unburned fuel and air mixture independently after the initiation of flame propagation, the tabulated chemistry concept is adopted to describe the auto-ignition chemistry. The methodology for extracting tabulated parameters from the chemical kinetics calculations is developed so that both cool flame reactions and main auto-ignition combustion can be well captured under a wider range of thermodynamic conditions. The SI-CAI hybrid combustion model (HCM) is then applied in the three-dimensional computational fluid dynamics (3-D CFD) engine simulation. The simulation results are compared with the experimental data obtained from a single cylinder VVA engine. The detailed analysis of the simulations demonstrates that the SI-CAI hybrid combustion process is characterised with the early flame propagation and subsequent multi-site auto-ignition around the main flame front, which is consistent with the optical results reported by other researchers. Besides, the systematic study of the in-cylinder condition reveals the influence mechanism of the early flame propagation on the subsequent auto-ignition.

  16. Maxi CAI with a Micro.

    ERIC Educational Resources Information Center

    Gerhold, George; And Others

    This paper describes an effective microprocessor-based CAI system which has been repeatedly tested by a large number of students and edited accordingly. Tasks not suitable for microprocessor based systems (authoring, testing, and debugging) were handled on larger multi-terminal systems. This approach requires that the CAI language used on the…

  17. CAI at CSDF: Organizational Strategies.

    ERIC Educational Resources Information Center

    Irwin, Margaret G.

    1982-01-01

    The computer assisted instruction (CAI) program at the California School for the Deaf, at Fremont, features individual Apple computers in classrooms as well as in CAI labs. When the whole class uses computers simultaneously, the teacher can help individuals, identify group weaknesses, note needs of the materials, and help develop additional CAI…

  18. A risk management approach to CAIS development

    NASA Technical Reports Server (NTRS)

    Hart, Hal; Kerner, Judy; Alden, Tony; Belz, Frank; Tadman, Frank

    1986-01-01

    The proposed DoD standard Common APSE Interface Set (CAIS) was developed as a framework set of interfaces that will support the transportability and interoperability of tools in the support environments of the future. While the current CAIS version is a promising start toward fulfilling those goals and current prototypes provide adequate testbeds for investigations in support of completing specifications for a full CAIS, there are many reasons why the proposed CAIS might fail to become a usable product and the foundation of next-generation (1990'S) project support environments such as NASA's Space Station software support environment. The most critical threats to the viability and acceptance of the CAIS include performance issues (especially in piggybacked implementations), transportability, and security requirements. To make the situation worse, the solution to some of these threats appears to be at conflict with the solutions to others.

  19. Coordinated Oxygen Isotopic and Petrologic Studies of CAIS Record Varying Composition of Protosolar

    NASA Technical Reports Server (NTRS)

    Simon, Justin I.; Matzel, J. E. P.; Simon, S. B.; Weber, P. K.; Grossman, L.; Ross, D. K.; Hutcheon, I. D.

    2012-01-01

    Ca-, Al-rich inclusions (CAIs) record the O-isotope composition of Solar nebular gas from which they grew [1]. High spatial resolution O-isotope measurements afforded by ion microprobe analysis across the rims and margin of CAIs reveal systematic variations in (Delta)O-17 and suggest formation from a diversity of nebular environments [2-4]. This heterogeneity has been explained by isotopic mixing between the O-16-rich Solar reservoir [6] and a second O-16-poor reservoir (probably nebular gas) with a "planetary-like" isotopic composition [e.g., 1, 6-7], but the mechanism and location(s) where these events occur within the protoplanetary disk remain uncertain. The orientation of large and systematic variations in (Delta)O-17 reported by [3] for a compact Type A CAI from the Efremovka reduced CV3 chondrite differs dramatically from reports by [4] of a similar CAI, A37 from the Allende oxidized CV3 chondrite. Both studies conclude that CAIs were exposed to distinct, nebular O-isotope reservoirs, implying the transfer of CAIs among different settings within the protoplanetary disk [4]. To test this hypothesis further and the extent of intra-CAI O-isotopic variation, a pristine compact Type A CAI, Ef-1 from Efremovka, and a Type B2 CAI, TS4 from Allende were studied. Our new results are equally intriguing because, collectively, O-isotopic zoning patterns in the CAIs indicate a progressive and cyclic record. The results imply that CAIs were commonly exposed to multiple environments of distinct gas during their formation. Numerical models help constrain conditions and duration of these events.

  20. CO-Bridged H-Cluster Intermediates in the Catalytic Mechanism of [FeFe]-Hydrogenase CaI

    DOE PAGES

    Ratzloff, Michael W.; Artz, Jacob H.; Mulder, David W.; ...

    2018-05-23

    The [FeFe]-hydrogenases ([FeFe] H 2ases) catalyze reversible H 2 activation at the H-cluster, which is composed of a [4Fe-4S] H subsite linked by a cysteine thiolate to a bridged, organometallic [2Fe-2S] ([2Fe] H) subsite. Profoundly different geometric models of the H-cluster redox states that orchestrate the electron/proton transfer steps of H 2 bond activation have been proposed. We have examined this question in the [FeFe] H 2ase I from Clostridium acetobutylicum (CaI) by Fourier-transform infrared (FTIR) spectroscopy with temperature annealing and H/D isotope exchange to identify the relevant redox states and define catalytic transitions. One-electron reduction of H ox ledmore » to formation of H redH + ([4Fe-4S] H 2+-Fe I-Fe I) and H red' ([4Fe-4S] H 1+-Fe II-Fe I), with both states characterized by low frequency μ-CO IR modes consistent with a fully bridged [2Fe] H. Similar μ-CO IR modes were also identified for H redH + of the [FeFe] H 2ase from Chlamydomonas reinhardtii (CrHydA1). The CaI proton-transfer variant C298S showed enrichment of an H/D isotope-sensitive μ-CO mode, a component of the hydride bound H-cluster IR signal, H hyd. Equilibrating CaI with increasing amounts of NaDT, and probed at cryogenic temperatures, showed H redH + was converted to H hyd. Over an increasing temperature range from 10 to 260 K catalytic turnover led to loss of Hhyd and appearance of H ox, consistent with enzymatic turnover and H 2 formation. The results show for CaI that the μ-CO of [2Fe] H remains bridging for all of the 'H red' states and that H redH + is on pathway to H hyd and H 2 evolution in the catalytic mechanism. Here, this provides a blueprint for designing small molecule catalytic analogs« less

  1. CO-Bridged H-Cluster Intermediates in the Catalytic Mechanism of [FeFe]-Hydrogenase CaI

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

    Ratzloff, Michael W.; Artz, Jacob H.; Mulder, David W.

    The [FeFe]-hydrogenases ([FeFe] H 2ases) catalyze reversible H 2 activation at the H-cluster, which is composed of a [4Fe-4S] H subsite linked by a cysteine thiolate to a bridged, organometallic [2Fe-2S] ([2Fe] H) subsite. Profoundly different geometric models of the H-cluster redox states that orchestrate the electron/proton transfer steps of H 2 bond activation have been proposed. We have examined this question in the [FeFe] H 2ase I from Clostridium acetobutylicum (CaI) by Fourier-transform infrared (FTIR) spectroscopy with temperature annealing and H/D isotope exchange to identify the relevant redox states and define catalytic transitions. One-electron reduction of H ox ledmore » to formation of H redH + ([4Fe-4S] H 2+-Fe I-Fe I) and H red' ([4Fe-4S] H 1+-Fe II-Fe I), with both states characterized by low frequency μ-CO IR modes consistent with a fully bridged [2Fe] H. Similar μ-CO IR modes were also identified for H redH + of the [FeFe] H 2ase from Chlamydomonas reinhardtii (CrHydA1). The CaI proton-transfer variant C298S showed enrichment of an H/D isotope-sensitive μ-CO mode, a component of the hydride bound H-cluster IR signal, H hyd. Equilibrating CaI with increasing amounts of NaDT, and probed at cryogenic temperatures, showed H redH + was converted to H hyd. Over an increasing temperature range from 10 to 260 K catalytic turnover led to loss of Hhyd and appearance of H ox, consistent with enzymatic turnover and H 2 formation. The results show for CaI that the μ-CO of [2Fe] H remains bridging for all of the 'H red' states and that H redH + is on pathway to H hyd and H 2 evolution in the catalytic mechanism. Here, this provides a blueprint for designing small molecule catalytic analogs« less

  2. CO-Bridged H-Cluster Intermediates in the Catalytic Mechanism of [FeFe]-Hydrogenase CaI.

    PubMed

    Ratzloff, Michael W; Artz, Jacob H; Mulder, David W; Collins, Reuben T; Furtak, Thomas E; King, Paul W

    2018-06-20

    The [FeFe]-hydrogenases ([FeFe] H 2 ases) catalyze reversible H 2 activation at the H-cluster, which is composed of a [4Fe-4S] H subsite linked by a cysteine thiolate to a bridged, organometallic [2Fe-2S] ([2Fe] H ) subsite. Profoundly different geometric models of the H-cluster redox states that orchestrate the electron/proton transfer steps of H 2 bond activation have been proposed. We have examined this question in the [FeFe] H 2 ase I from Clostridium acetobutylicum (CaI) by Fourier-transform infrared (FTIR) spectroscopy with temperature annealing and H/D isotope exchange to identify the relevant redox states and define catalytic transitions. One-electron reduction of H ox led to formation of H red H + ([4Fe-4S] H 2+ -Fe I -Fe I ) and H red ' ([4Fe-4S] H 1+ -Fe II -Fe I ), with both states characterized by low frequency μ-CO IR modes consistent with a fully bridged [2Fe] H . Similar μ-CO IR modes were also identified for H red H + of the [FeFe] H 2 ase from Chlamydomonas reinhardtii (CrHydA1). The CaI proton-transfer variant C298S showed enrichment of an H/D isotope-sensitive μ-CO mode, a component of the hydride bound H-cluster IR signal, H hyd . Equilibrating CaI with increasing amounts of NaDT, and probed at cryogenic temperatures, showed H red H + was converted to H hyd . Over an increasing temperature range from 10 to 260 K catalytic turnover led to loss of H hyd and appearance of H ox , consistent with enzymatic turnover and H 2 formation. The results show for CaI that the μ-CO of [2Fe] H remains bridging for all of the "H red " states and that H red H + is on pathway to H hyd and H 2 evolution in the catalytic mechanism. These results provide a blueprint for designing small molecule catalytic analogs.

  3. The Vibrio cholerae quorum-sensing autoinducer CAI-1: analysis of the biosynthetic enzyme CqsA

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

    Kelly, R.; Bolitho, M; Higgins, D

    2009-01-01

    Vibrio cholerae, the bacterium that causes the disease cholera, controls virulence factor production and biofilm development in response to two extracellular quorum-sensing molecules, called autoinducers. The strongest autoinducer, called CAI-1 (for cholera autoinducer-1), was previously identified as (S)-3-hydroxytridecan-4-one. Biosynthesis of CAI-1 requires the enzyme CqsA. Here, we determine the CqsA reaction mechanism, identify the CqsA substrates as (S)-2-aminobutyrate and decanoyl coenzyme A, and demonstrate that the product of the reaction is 3-aminotridecan-4-one, dubbed amino-CAI-1. CqsA produces amino-CAI-1 by a pyridoxal phosphate-dependent acyl-CoA transferase reaction. Amino-CAI-1 is converted to CAI-1 in a subsequent step via a CqsA-independent mechanism. Consistent with this,more » we find cells release {ge}100 times more CAI-1 than amino-CAI-1. Nonetheless, V. cholerae responds to amino-CAI-1 as well as CAI-1, whereas other CAI-1 variants do not elicit a quorum-sensing response. Thus, both CAI-1 and amino-CAI-1 have potential as lead molecules in the development of an anticholera treatment.« less

  4. Oxygen Isotope Measurements of a Rare Murchison Type A CAI and Its Rim

    NASA Technical Reports Server (NTRS)

    Matzel, J. E. P.; Simon, J. I.; Hutcheon, I. D.; Jacobsen, B.; Simon, S. B.; Grossman, L.

    2013-01-01

    Ca-, Al-rich inclusions (CAIs) from CV chondrites commonly show oxygen isotope heterogeneity among different mineral phases within individual inclusions reflecting the complex history of CAIs in both the solar nebula and/or parent bodies. The degree of isotopic exchange is typically mineral-specific, yielding O-16-rich spinel, hibonite and pyroxene and O-16-depleted melilite and anorthite. Recent work demonstrated large and systematic variations in oxygen isotope composition within the margin and Wark-Lovering rim of an Allende Type A CAI. These variations suggest that some CV CAIs formed from several oxygen reservoirs and may reflect transport between distinct regions of the solar nebula or varying gas composition near the proto-Sun. Oxygen isotope compositions of CAIs from other, less-altered chondrites show less intra-CAI variability and 16O-rich compositions. The record of intra-CAI oxygen isotope variability in CM chondrites, which commonly show evidence for low-temperature aqueous alteration, is less clear, in part because the most common CAIs found in CM chondrites are mineralogically simple (hibonite +/- spinel or spinel +/- pyroxene) and are composed of minerals less susceptible to O-isotopic exchange. No measurements of the oxygen isotope compositions of rims on CAIs in CM chondrites have been reported. Here, we present oxygen isotope data from a rare, Type A CAI from the Murchison meteorite, MUM-1. The data were collected from melilite, hibonite, perovskite and spinel in a traverse into the interior of the CAI and from pyroxene, melilite, anorthite, and spinel in the Wark-Lovering rim. Our objectives were to (1) document any evidence for intra-CAI oxygen isotope variability; (2) determine the isotopic composition of the rim minerals and compare their composition(s) to the CAI interior; and (3) compare the MUM-1 data to oxygen isotope zoning profiles measured from CAIs in other chondrites.

  5. Fine-Gained CAIs in Comet Samples: Moderate Refractory Character and Comparison to Small Refractory Inclusions in Chondrites

    NASA Technical Reports Server (NTRS)

    Joswiak, D. J.; Brownlee, D. E.; Nguyen, A. N.; Messenger, S

    2017-01-01

    Examination of >200 comet Wild 2 particles collected by the Stardust (SD) mission shows that the CAI abundance of comet Wild 2's rocky material is near 1% and that nearly 50% of all bulbous tracks will contain at least one recognizable CAI fragment. A similar abundance to Wild 2 is found in a giant cluster IDP thought to be of cometary origin. The properties of these CAIs and their comparison with meteoritic CAIs provide important clues on the role of CAIs in the early Solar System (SS) and how they were transported to the edge of the solar nebula where Kuiper Belt comets formed. Previously, only two CAIs in comet Wild 2 had been identified and studied in detail. Here we present 2 new Wild 2 CAIs and 2 from a giant cluster cometary IDP, describe their mineralogical characteristics and show that they are most analogous to nodules in spinel-rich, fine-grained inclusions (FGIs) observed in CV3 and other chondrites. Additionally, we present new O isotope measurements from one CAI from comet Wild 2 and show that its oxygen isotopic composition is similar to some FGIs. This is only the second CAI from Wild 2 in which O isotopes have been measured.

  6. The Screen Display Syntax for CAI.

    ERIC Educational Resources Information Center

    Richards, Boyd F.; Salisbury, David F.

    1987-01-01

    Describes four storyboard techniques frequently used in designing computer assisted instruction (CAI) programs, and explains screen display syntax (SDS), a new technique combining the major advantages of the storyboard techniques. SDS was developed to facilitate communication among designers, programmers, and editors working on a large CAI basic…

  7. CAI: Its Cost and Its Role.

    ERIC Educational Resources Information Center

    Pressman, Israel; Rosenbloom, Bruce

    1984-01-01

    Describes and evaluates costs of hardware, software, training, and maintenance for computer assisted instruction (CAI) as they relate to total system cost. An example of an educational system provides an illustration of CAI cost analysis. Future developments, cost effectiveness, affordability, and applications in public and private environments…

  8. Quantum Computational Universality of the 2D Cai-Miyake-D"ur-Briegel Quantum State

    NASA Astrophysics Data System (ADS)

    Wei, Tzu-Chieh; Raussendorf, Robert; Kwek, Leong Chuan

    2012-02-01

    Universal quantum computation can be achieved by simply performing single-qubit measurements on a highly entangled resource state, such as cluster states. Cai, Miyake, D"ur, and Briegel recently constructed a ground state of a two-dimensional quantum magnet by combining multiple Affleck-Kennedy-Lieb-Tasaki quasichains of mixed spin-3/2 and spin-1/2 entities and by mapping pairs of neighboring spin-1/2 particles to individual spin-3/2 particles [Phys. Rev. A 82, 052309 (2010)]. They showed that this state enables universal quantum computation by constructing single- and two-qubit universal gates. Here, we give an alternative understanding of how this state gives rise to universal measurement-based quantum computation: by local operations, each quasichain can be converted to a one-dimensional cluster state and entangling gates between two neighboring logical qubits can be implemented by single-spin measurements. Furthermore, a two-dimensional cluster state can be distilled from the Cai-Miyake-D"ur-Briegel state.

  9. Search for 41K Excess in Efremovka CAIs

    NASA Astrophysics Data System (ADS)

    Srinivasan, G.; Ulyanov, A. A.; Goswami, J. N.

    1993-07-01

    We have used the ion microprobe to measure K isotopic composition of refractory phases in Efremovka CAIs to look for the possible presence of K excess from the decay of extinct radionuclide Ca (halflife = 0.13 Ma). The presence of Ca at the time of CAI formation, if established, will allow us to place a lower limit on the time interval between the last injection of freshly synthesized matter into the solar nebula and the formation of some of the first solid objects (CAIs) in the solar system. Several attempts have been made earlier to detect 41K excess in Allende CAIs [1-4]. We have further investigated this problem by analyzing the Efremovka CAIs for two reasons. First, both the petrographic and magnesium isotopic systematics suggest the Efremovka CAIs to be less altered compared to the Allende CAIs making them an ideal and perhaps better sample for this study. Second, the presence of large perovskite (~10 micrometers) allowed us to analyse this phase, which was not included in earlier studies. The major difficulty in accurately measuring 41K, which was identified in earlier studies, is the unresolvable (40Ca42Ca)++ interference, which was found to be matrix dependent [4]. In addition, one can also have interfernce from the (40CaH)+ peak. In our operating condition the interference from the hydride peak can be neglected (Fig. 1, which appears in the hard copy). We have analyzed terrestrial perovskite (K <= 20 ppm) to determine the (40Ca42Ca)++ correction term, and its equivalence with (40Ca43Ca)++ ion signal at mass 41.5 [4]. In perovskite, the (40Ca42Ca)++ signal constitutes ~80% of the signal at 41K and we could estimate this interference with confidence. A value of (2.7 +- 0.1) x 10^-5 was obtained for the ratio [(40Ca42Ca)++/42Ca+], which is similar to the measured [(40Ca43Ca)++/43Ca+] ratio of (2.4 +- 0.2) x 10^-5. We have therefore used the measured value for the latter ratio in the analyzed phases to correct for the doubly charged interference at mass 41

  10. Implications of Windowing Techniques for CAI.

    ERIC Educational Resources Information Center

    Heines, Jesse M.; Grinstein, Georges G.

    This paper discusses the use of a technique called windowing in computer assisted instruction to allow independent control of functional areas in complex CAI displays and simultaneous display of output from a running computer program and coordinated instructional material. Two obstacles to widespread use of CAI in computer science courses are…

  11. Personality preference influences medical student use of specific computer-aided instruction (CAI)

    PubMed Central

    McNulty, John A; Espiritu, Baltazar; Halsey, Martha; Mendez, Michelle

    2006-01-01

    Background The objective of this study was to test the hypothesis that personality preference, which can be related to learning style, influences individual utilization of CAI applications developed specifically for the undergraduate medical curriculum. Methods Personality preferences of students were obtained using the Myers-Briggs Type Indicator (MBTI) test. CAI utilization for individual students was collected from entry logs for two different web-based applications (a discussion forum and a tutorial) used in the basic science course on human anatomy. Individual login data were sorted by personality preference and the data statistically analyzed by 2-way mixed ANOVA and correlation. Results There was a wide discrepancy in the level and pattern of student use of both CAI. Although individual use of both CAI was positively correlated irrespective of MBTI preference, students with a "Sensing" preference tended to use both CAI applications more than the "iNtuitives". Differences in the level of use of these CAI applications (i.e., higher use of discussion forum vs. a tutorial) were also found for the "Perceiving/Judging" dimension. Conclusion We conclude that personality/learning preferences of individual students influence their use of CAI in the medical curriculum. PMID:16451719

  12. Design specifications for NALDA (Naval Aviation Logistics Data Analysis) CAI (computer aided instruction): Phase 2, Interim report

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

    Twitty, A.F.; Handler, B.H.; Duncan, L.D.

    Data Systems Engineering Organization (DSEO) personnel are developing a prototype computer aided instruction (CAI) system for the Naval Aviation Logistics Data Analysis (NALDA) system. The objective of this project is to provide a prototype for implementing CAI as an enhancement to existing NALDA training. The CAI prototype project is being performed in phases. The task undertaken in Phase I was to analyze the problem and the alternative solutions and to develop a set of recommendations on how best to proceed. In Phase II a structured design and specification document was completed that will provide the basis for development and implementationmore » of the desired CAI system. Phase III will consist of designing, developing, and testing a user interface which will extend the features of the Phase II prototype. The design of the CAI prototype has followed a rigorous structured analysis based on Yourdon/DeMarco methodology and Information Engineering tools. This document includes data flow diagrams, a data dictionary, process specifications, an entity-relationship diagram, a curriculum description, special function key definitions, and a set of standards developed for the NALDA CAI Prototype.« less

  13. NALDA (Naval Aviation Logistics Data Analysis) CAI (computer aided instruction)

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

    Handler, B.H.; France, P.A.; Frey, S.C.

    Data Systems Engineering Organization (DSEO) personnel developed a prototype computer aided instruction CAI system for the Naval Aviation Logistics Data Analysis (NALDA) system. The objective of this project was to provide a CAI prototype that could be used as an enhancement to existing NALDA training. The CAI prototype project was performed in phases. The task undertaken in Phase I was to analyze the problem and the alternative solutions and to develop a set of recommendations on how best to proceed. The findings from Phase I are documented in Recommended CAI Approach for the NALDA System (Duncan et al., 1987). Inmore » Phase II, a structured design and specifications were developed, and a prototype CAI system was created. A report, NALDA CAI Prototype: Phase II Final Report, was written to record the findings and results of Phase II. NALDA CAI: Recommendations for an Advanced Instructional Model, is comprised of related papers encompassing research on computer aided instruction CAI, newly developing training technologies, instructional systems development, and an Advanced Instructional Model. These topics were selected because of their relevancy to the CAI needs of NALDA. These papers provide general background information on various aspects of CAI and give a broad overview of new technologies and their impact on the future design and development of training programs. The paper within have been index separately elsewhere.« less

  14. The Evolutionary Development of CAI Hardware.

    ERIC Educational Resources Information Center

    Stifle, John E.

    After six years of research in computer assisted instruction (CAI) using PLATO III, a decision was made at the University of Illinois to develop a larger system as a national CAI resource. This document describes the design specifications and problems in the development of PLATO IV, a system which is capable of accomodating up to 4,000 terminals…

  15. Creation and Distribution of CAIs in the Protoplanetary Nebula

    NASA Technical Reports Server (NTRS)

    Cuzzi, J. N.; Davis, S. S.; Dobrovolskis, A. R.

    2003-01-01

    CaAl rich refractory mineral inclusions (CAIs) found at 1 - 10% mass fraction in primitive chondrites appear to be several million years older than the dominant (chondrule) components in the same parent bodies. A prevalent concern is that it is difficult to retain CAIs for this long against gas-drag-induced radial drift into the sun. We assess a hot inner (turbulent) nebula context for CAI formation, using analytical models of nebula evolution and particle diffusion. We show that outward radial diffusion in a weakly turbulent nebula can prevent significant numbers of CAI-size particles from being lost into the sun for times of 1 - 3 x 10(exp 6) years. To match the CAI abundances quantitatively, we advocate an enhancement of the inner hot nebula in silicate-forming material, due to rapid inward migration of very primitive, silicate and carbon rich, meter-sized objects. 'Combustion' of the carbon into CO would make the CAI formation environment more reduced than solar, as certain observations imply. Abundant CO might also play a role in mass-independent chemical fractionation of oxygen isotopes as seen in CAIs and associated primitive, high-temperature condensates.

  16. Computers for Your Classroom: CAI and CMI.

    ERIC Educational Resources Information Center

    Thomas, David B.; Bozeman, William C.

    1981-01-01

    The availability of compact, low-cost computer systems provides a means of assisting classroom teachers in the performance of their duties. Computer-assisted instruction (CAI) and computer-managed instruction (CMI) are two applications of computer technology with which school administrators should become familiar. CAI is a teaching medium in which…

  17. Research on TRIZ and CAIs Application Problems for Technology Innovation

    NASA Astrophysics Data System (ADS)

    Li, Xiangdong; Li, Qinghai; Bai, Zhonghang; Geng, Lixiao

    In order to realize application of invent problem solve theory (TRIZ) and computer aided innovation software (CAIs) , need to solve some key problems, such as the mode choice of technology innovation, establishment of technology innovation organization network(TION), and achievement of innovative process based on TRIZ and CAIs, etc.. This paper shows that the demands for TRIZ and CAIs according to the characteristics and existing problem of the manufacturing enterprises. Have explained that the manufacturing enterprises need to set up an open TION of enterprise leading type, and achieve the longitudinal cooperation innovation with institution of higher learning. The process of technology innovation based on TRIZ and CAIs has been set up from researching and developing point of view. Application of TRIZ and CAIs in FY Company has been summarized. The application effect of TRIZ and CAIs has been explained using technology innovation of the close goggle valve product.

  18. CAIs in Semarkona (LL3.0)

    NASA Technical Reports Server (NTRS)

    Mishra, R. K.; Simon, J. I.; Ross, D. K.; Marhas, K. K.

    2016-01-01

    Calcium, Aluminum-rich inclusions (CAIs) are the first forming solids of the Solar system. Their observed abundance, mean size, and mineralogy vary quite significantly between different groups of chondrites. These differences may reflect the dynamics and distinct cosmochemical conditions present in the region(s) of the protoplanetary disk from which each type likely accreted. Only about 11 such objects have been found in L and LL type while another 57 have been found in H type ordinary chondrites, compared to thousands in carbonaceous chondrites. At issue is whether the rare CAIs contained in ordinary chondrites truly reflect a distinct population from the inclusions commonly found in other chondrite types. Semarkona (LL3.00) (fall, 691 g) is the most pristine chondrite available in our meteorite collection. Here we report petrography and mineralogy of 3 CAIs from Semarkona

  19. The Effect of CAI on Reading Achievement.

    ERIC Educational Resources Information Center

    Hardman, Regina

    A study determined whether computer assisted instruction (CAI) had an effect on students' reading achievement. Subjects were 21 randomly selected fourth-grade students at D. S. Wentworth Elementary School on the south side of Chicago in a low-income neighborhood who received a year's exposure to a CAI program, and 21 randomly selected students at…

  20. Individual Differences in Learner Controlled CAI.

    ERIC Educational Resources Information Center

    Judd, Wilson A.; And Others

    Two assumptions in support of learner-controlled computer-assisted instruction (CAI) are that (1) instruction administered under learner control will be less aversive than if administered under program control, and (2) the student is sufficiently aware of his learning state to make, in most instances, his own instructional decisions. Some 130…

  1. Evaluation Criteria for Micro-CAI: A Psychometric Approach

    PubMed Central

    Wallace, Douglas; Slichter, Mark; Bolwell, Christine

    1985-01-01

    The increased use of microcomputer-based instructional programs has resulted in a greater need for third-party evaluation of the software. This in turn has prompted the development of micro-CAI evaluation tools. The present project sought to develop a prototype instrument to assess the impact of CAI program presentation characteristics on students. Data analysis and scale construction was conducted using standard item reliability analyses and factor analytic techniques. Adequate subscale reliabilities and factor structures were found, suggesting that a psychometric approach to CAI evaluation may possess some merit. Efforts to assess the utility of the resultant instrument are currently underway.

  2. Compound ultrarefractory CAI-bearing inclusions from CV3 carbonaceous chondrites

    NASA Astrophysics Data System (ADS)

    Ivanova, Marina A.; Krot, Alexander N.; Nagashima, Kazuhide; MacPherson, Glenn J.

    2012-12-01

    Abstract-Two compound calcium-aluminum-rich inclusions (<span class="hlt">CAIs</span>), 3N from the oxidized CV chondrite Northwest Africa (NWA) 3118 and 33E from the reduced CV chondrite Efremovka, contain ultrarefractory (UR) inclusions. 3N is a forsterite-bearing type B (FoB) <span class="hlt">CAI</span> that encloses UR inclusion 3N-24 composed of Zr,Sc,Y-rich oxides, Y-rich perovskite, and Zr,Sc-rich Al,Ti-diopside. 33E contains a fluffy type A (FTA) <span class="hlt">CAI</span> and UR <span class="hlt">CAI</span> 33E-1, surrounded by Wark-Lovering rim layers of spinel, Al-diopside, and forsterite, and a common forsterite-rich accretionary rim. 33E-1 is composed of Zr,Sc,Y-rich oxides, Y-rich perovskite, Zr,Sc,Y-rich pyroxenes (Al,Ti-diopside, Sc-rich pyroxene), and gehlenite. 3N-24's UR oxides and Zr,Sc-rich Al,Ti-diopsides are 16O-poor (Δ17O approximately -<span class="hlt">2</span>‰ to -5‰). Spinel in 3N-24 and spinel and Al-diopside in the FoB <span class="hlt">CAI</span> are 16O-rich (Δ17O approximately -23 ± <span class="hlt">2</span>‰). 33E-1's UR oxides and Zr,Sc-rich Al,Ti-diopsides are 16O-depleted (Δ17O approximately -<span class="hlt">2</span>‰ to -5‰) vs. Al,Ti-diopside of the FTA <span class="hlt">CAI</span> and spinel (Δ17O approximately -23 ± <span class="hlt">2</span>‰), and Wark-Lovering rim Al,Ti-diopside (Δ17O approximately -7‰ to -19‰). We infer that the inclusions experienced multistage formation in nebular regions with different oxygen-isotope compositions. 3N-24 and 33E-1's precursors formed by evaporation/condensation above 1600 °C. 3N and 33E's precursors formed by condensation and melting (3N only) at significantly lower temperatures. 3N-24 and 3N's precursors aggregated into a compound object and experienced partial melting and thermal annealing. 33E-1 and 33E avoided melting prior to and after aggregation. They acquired Wark-Lovering and common forsterite-rich accretionary rims, probably by condensation, followed by thermal annealing. We suggest 3N-24 and 33E-1 originated in a 16O-rich gaseous reservoir and subsequently experienced isotope exchange in a 16O-poor gaseous reservoir. Mechanism and timing of oxygen-isotope exchange remain</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018GeCoA.221..296K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018GeCoA.221..296K"><span>A multielement isotopic study of refractory FUN and F <span class="hlt">CAIs</span>: Mass-dependent and mass-independent isotope effects</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kööp, Levke; Nakashima, Daisuke; Heck, Philipp R.; Kita, Noriko T.; Tenner, Travis J.; Krot, Alexander N.; Nagashima, Kazuhide; Park, Changkun; Davis, Andrew M.</p> <p>2018-01-01</p> <p>Calcium-aluminum-rich inclusions (<span class="hlt">CAIs</span>) are the oldest dated objects that formed inside the Solar System. Among these are rare, enigmatic objects with large mass-dependent fractionation effects (F <span class="hlt">CAIs</span>), which sometimes also have large nucleosynthetic anomalies and a low initial abundance of the short-lived radionuclide 26Al (FUN <span class="hlt">CAIs</span>). We have studied seven refractory hibonite-rich <span class="hlt">CAIs</span> and one grossite-rich <span class="hlt">CAI</span> from the Murchison (CM<span class="hlt">2</span>) meteorite for their oxygen, calcium, and titanium isotopic compositions. The 26Al-26Mg system was also studied in seven of these <span class="hlt">CAIs</span>. We found mass-dependent heavy isotope enrichment in all measured elements, but never simultaneously in the same <span class="hlt">CAI</span>. The data are hard to reconcile with a single-stage melt evaporation origin and may require reintroduction or reequilibration for magnesium, oxygen and titanium after evaporation for some of the studied <span class="hlt">CAIs</span>. The initial 26Al/27Al ratios inferred from model isochrons span a range from <1 × 10-6 to canonical (∼5 × 10-5). The <span class="hlt">CAIs</span> show a mutual exclusivity relationship between inferred incorporation of live 26Al and the presence of resolvable anomalies in 48Ca and 50Ti. Furthermore, a relationship exists between 26Al incorporation and Δ17O in the hibonite-rich <span class="hlt">CAIs</span> (i.e., 26Al-free <span class="hlt">CAIs</span> have resolved variations in Δ17O, while <span class="hlt">CAIs</span> with resolved 26Mg excesses have Δ17O values close to -23‰). Only the grossite-rich <span class="hlt">CAI</span> has a relatively enhanced Δ17O value (∼-17‰) in spite of a near-canonical 26Al/27Al. We interpret these data as indicating that fractionated hibonite-rich <span class="hlt">CAIs</span> formed over an extended time period and sampled multiple stages in the isotopic evolution of the solar nebula, including: (1) an 26Al-poor nebula with large positive and negative anomalies in 48Ca and 50Ti and variable Δ17O; (<span class="hlt">2</span>) a stage of 26Al-admixture, during which anomalies in 48Ca and 50Ti had been largely diluted and a Δ17O value of ∼-23‰ had been achieved in the <span class="hlt">CAI</span> formation region; and (3</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/965072','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/965072"><span>Silicon Isotopic Fractionation of <span class="hlt">CAI</span>-like Vacuum Evaporation Residues</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>Knight, K; Kita, N; Mendybaev, R</p> <p>2009-06-18</p> <p>Calcium-, aluminum-rich inclusions (<span class="hlt">CAIs</span>) are often enriched in the heavy isotopes of magnesium and silicon relative to bulk solar system materials. It is likely that these isotopic enrichments resulted from evaporative mass loss of magnesium and silicon from early solar system condensates while they were molten during one or more high-temperature reheating events. Quantitative interpretation of these enrichments requires laboratory determinations of the evaporation kinetics and associated isotopic fractionation effects for these elements. The experimental data for the kinetics of evaporation of magnesium and silicon and the evaporative isotopic fractionation of magnesium is reasonably complete for Type B <span class="hlt">CAI</span> liquidsmore » (Richter et al., 2002, 2007a). However, the isotopic fractionation factor for silicon evaporating from such liquids has not been as extensively studied. Here we report new ion microprobe silicon isotopic measurements of residual glass from partial evaporation of Type B <span class="hlt">CAI</span> liquids into vacuum. The silicon isotopic fractionation is reported as a kinetic fractionation factor, {alpha}{sub Si}, corresponding to the ratio of the silicon isotopic composition of the evaporation flux to that of the residual silicate liquid. For <span class="hlt">CAI</span>-like melts, we find that {alpha}{sub Si} = 0.98985 {+-} 0.00044 (<span class="hlt">2</span>{sigma}) for {sup 29}Si/{sup 28}Si with no resolvable variation with temperature over the temperature range of the experiments, 1600-1900 C. This value is different from what has been reported for evaporation of liquid Mg{sub <span class="hlt">2</span>}SiO{sub 4} (Davis et al., 1990) and of a melt with CI chondritic proportions of the major elements (Wang et al., 2001). There appears to be some compositional control on {alpha}{sub Si}, whereas no compositional effects have been reported for {alpha}{sub Mg}. We use the values of {alpha}Si and {alpha}Mg, to calculate the chemical compositions of the unevaporated precursors of a number of isotopically fractionated <span class="hlt">CAIs</span> from CV chondrites</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012JIPM...29..667T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012JIPM...29..667T"><span><span class="hlt">CAI</span> System of Obunsha Co., Ltd. Using CD-ROM</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Todokoro, Shigeru; Mukai, Yoshihiro</p> <p></p> <p>This paper introduces the present status of R & D on <span class="hlt">CAI</span> teaching materials in Obunsha Co., Ltd. Characteristics of <span class="hlt">CAI</span> using CD-ROM as well as Culture-in <span class="hlt">CAI</span> Teaching Materials System for junior high school English are described. The system consists of CD-ROM driver XM-2000 and Pasopia 700 of Toshiba Corporation having both features of CD-ROM and FD. CD-ROM stores vast amount of voice data while FD does text and graphics data. It is a frame-oriented mode system enabling to raise learning effect.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22624244','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22624244"><span>Particulated articular cartilage: <span class="hlt">CAIS</span> and DeNovo NT.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Farr, Jack; Cole, Brian J; Sherman, Seth; Karas, Vasili</p> <p>2012-03-01</p> <p>Cartilage Autograft Implantation System (<span class="hlt">CAIS</span>; DePuy/Mitek, Raynham, MA) and DeNovo Natural Tissue (NT; ISTO, St. Louis, MO) are novel treatment options for focal articular cartilage defects in the knee. These methods involve the implantation of particulated articular cartilage from either autograft or juvenile allograft donor, respectively. In the laboratory and in animal models, both <span class="hlt">CAIS</span> and DeNovo NT have demonstrated the ability of the transplanted cartilage cells to "escape" from the extracellular matrix, migrate, multiply, and form a new hyaline-like cartilage tissue matrix that integrates with the surrounding host tissue. In clinical practice, the technique for both <span class="hlt">CAIS</span> and DeNovo NT is straightforward, requiring only a single surgery to affect cartilage repair. Clinical experience is limited, with short-term studies demonstrating both procedures to be safe, feasible, and effective, with improvements in subjective patient scores, and with magnetic resonance imaging evidence of good defect fill. While these treatment options appear promising, prospective randomized controlled studies are necessary to refine the indications and contraindications for both <span class="hlt">CAIS</span> and DeNovo NT.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018E%26PSL.482..324A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018E%26PSL.482..324A"><span>Closed system oxygen isotope redistribution in igneous <span class="hlt">CAIs</span> upon spinel dissolution</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Aléon, Jérôme</p> <p>2018-01-01</p> <p>In several Calcium-Aluminum-rich Inclusions (<span class="hlt">CAIs</span>) from the CV3 chondrites Allende and Efremovka, representative of the most common igneous <span class="hlt">CAI</span> types (type A, type B and Fractionated with Unknown Nuclear isotopic anomalies, FUN), the relationship between 16O-excesses and TiO<span class="hlt">2</span> content in pyroxene indicates that the latter commonly begins to crystallize with a near-terrestrial 16O-poor composition and becomes 16O-enriched during crystallization, reaching a near-solar composition. Mass balance calculations were performed to investigate the contribution of spinel to this 16O-enrichment. It is found that a back-reaction of early-crystallized 16O-rich spinel with a silicate partial melt having undergone a 16O-depletion is consistent with the O isotopic evolution of <span class="hlt">CAI</span> minerals during magmatic crystallization. Dissolution of spinel explains the O isotopic composition (16O-excess and extent of mass fractionation) of pyroxene as well as that of primary anorthite/dmisteinbergite and possibly that of the last melilite crystallizing immediately before pyroxene. It requires that igneous <span class="hlt">CAIs</span> behaved as closed-systems relative to oxygen from nebular gas during a significant fraction of their cooling history, contrary to the common assumption that <span class="hlt">CAI</span> partial melts constantly equilibrated with gas. The mineralogical control on O isotopes in igneous <span class="hlt">CAIs</span> is thus simply explained by a single 16O-depletion during magmatic crystallization. This 16O-depletion occurred in an early stage of the thermal history, after the crystallization of spinel, i.e. in the temperature range for melilite crystallization/partial melting and did not require multiple, complex or late isotope exchange. More experimental work is however required to deduce the protoplanetary disk conditions associated with this 16O-depletion.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20150002917','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20150002917"><span>Two Generations of Sodic Metasomatism in an Allende Type B <span class="hlt">CAI</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Ross, D. K.; Simon, J. I.; Simon, S. B.; Grossman, L.</p> <p>2015-01-01</p> <p>Calcium-Aluminum rich inclusions (<span class="hlt">CAI</span>) in Allende, along with other chondritic compo-nents, experienced variable amounts and types of alter-ation of their mineralogy and chemistry. In <span class="hlt">CAIs</span>, one of the principal types of alteration led to the depo-sition of nepheline and sodalite. Here we extend initial obervations of alteration in an Allende <span class="hlt">CAI</span>, focus-ing on occurences of nepheline and a nepheline-like phase with unusally high Ca (referred to as "calcic nepheline" in this abstract). Detailed petrographic and microchemical observations of alteration phases in an Allende Type B <span class="hlt">CAI</span> (TS4) show that two separate generations of "nepheline", with very distinct composi-tions, crystallized around the margins and in the interi-or of this <span class="hlt">CAI</span>. We use observations of micro-faults as potential temporal markers, in order to place constraints on the timing of alteration events in Allende. These observa-tions of micro-faulting that truncate and offset one gen-eration of "nepheline" indicate that some "nepheline" crystallized before incorporation of the <span class="hlt">CAI</span> into the Allende parent-body. Some of the sodic metasomatism in some Allende <span class="hlt">CAIs</span> occurred prior to Allende par-ent-body assembly. The earlier generation of "calcic-nepheline" has a very distinctive, calcium-rich compo-sition, and the second generation is low in calcium, and matches the compositions of nephelines found in near-by altered chondrules, and in the Allende matrix.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AGUFM.A41I0176K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFM.A41I0176K"><span>GOSAT CO<span class="hlt">2</span> retrieval results using TANSO-<span class="hlt">CAI</span> aerosol information over East Asia</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>KIM, M.; Kim, W.; Jung, Y.; Lee, S.; Kim, J.; Lee, H.; Boesch, H.; Goo, T. Y.</p> <p>2015-12-01</p> <p>In the satellite remote sensing of CO<span class="hlt">2</span>, incorrect aerosol information could induce large errors as previous studies suggested. Many factors, such as, aerosol type, wavelength dependency of AOD, aerosol polarization effect and etc. have been main error sources. Due to these aerosol effects, large number of data retrieved are screened out in quality control, or retrieval errors tend to increase if not screened out, especially in East Asia where aerosol concentrations are fairly high. To reduce these aerosol induced errors, a CO<span class="hlt">2</span> retrieval algorithm using the simultaneous TANSO-<span class="hlt">CAI</span> aerosol information is developed. This algorithm adopts AOD and aerosol type information as a priori information from the <span class="hlt">CAI</span> aerosol retrieval algorithm. The CO<span class="hlt">2</span> retrieval algorithm based on optimal estimation method and VLIDORT, a vector discrete ordinate radiative transfer model. The CO<span class="hlt">2</span> algorithm, developed with various state vectors to find accurate CO<span class="hlt">2</span> concentration, shows reasonable results when compared with other dataset. This study concentrates on the validation of retrieved results with the ground-based TCCON measurements in East Asia and the comparison with the previous retrieval from ACOS, NIES, and UoL. Although, the retrieved CO<span class="hlt">2</span> concentration is lower than previous results by ppm's, it shows similar trend and high correlation with previous results. Retrieved data and TCCON measurements data are compared at three stations of Tsukuba, Saga, Anmyeondo in East Asia, with the collocation criteria of ±<span class="hlt">2</span>°in latitude/longitude and ±1 hours of GOSAT passing time. Compared results also show similar trend with good correlation. Based on the TCCON comparison results, bias correction equation is calculated and applied to the East Asia data.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://eric.ed.gov/?q=binary+AND+number+AND+system+AND+computers&pg=3&id=ED198791','ERIC'); return false;" href="https://eric.ed.gov/?q=binary+AND+number+AND+system+AND+computers&pg=3&id=ED198791"><span>A Multi-Media <span class="hlt">CAI</span> Terminal Based upon a Microprocessor with Applications for the Handicapped.</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>Brebner, Ann; Hallworth, H. J.</p> <p></p> <p>The design of the <span class="hlt">CAI</span> interface described is based on the microprocessor in order to meet three basic requirements for providing appropriate instruction to the developmentally handicapped: (1) portability, so that <span class="hlt">CAI</span> can be taken into the customary learning environment; (<span class="hlt">2</span>) reliability; and (3) flexibility, to permit use of new input and output…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3065330','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3065330"><span>ROS-activated Ca/calmodulin kinase IIδ is required for late INa augmentation leading to cellular <span class="hlt">Na</span> and Ca overload</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Wagner, Stefan; Ruff, Hanna M.; Weber, Sarah L.; Bellmann, Sarah; Sowa, Thomas; Schulte, Timo; Grandi, Eleonora; Bers, Donald M.; Backs, Johannes; Belardinelli, Luiz; Maier, Lars S.</p> <p>2011-01-01</p> <p>Rationale In heart failure (HF), CaMKII expression and reactive oxygen species (ROS) are increased. Both ROS and CaMKII can increase late INa leading to intracellular <span class="hlt">Na</span> accumulation and arrhythmias. It has been shown that ROS can activate CaMKII via oxidation. Objective We tested whether CaMKIIδ is required for ROS-dependent late INa regulation and if ROS-induced Ca released from the sarcoplasmic reticulum (SR) is involved. Methods and Results 40 µmol/L H<span class="hlt">2</span>O<span class="hlt">2</span> significantly increased CaMKII oxidation and autophosphorylation in permeabilized rabbit cardiomyocytes. Without free [<span class="hlt">Ca]i</span> (5 mmol/L BAPTA/1 mmol/L Br<span class="hlt">2</span>-BAPTA) or after SR depletion (caffeine 10 mmol/L, thapsigargin 5 µmol/L) the H<span class="hlt">2</span>O<span class="hlt">2</span>-dependent CaMKII oxidation and autophosphorylation was abolished. H<span class="hlt">2</span>O<span class="hlt">2</span> significantly increased SR Ca spark frequency (confocal microscopy) but reduced SR Ca load. In wildtype (WT) mouse myocytes, H<span class="hlt">2</span>O<span class="hlt">2</span> increased late INa (whole cell patch-clamp). This increase was abolished in CaMKIIδ−/− myocytes. H<span class="hlt">2</span>O<span class="hlt">2</span>-induced [<span class="hlt">Na</span>]i and [<span class="hlt">Ca]i</span> accumulation (SBFI and Indo-1 epifluorescence) was significantly slowed in CaMKIIδ−/− myocytes (vs. WT). CaMKIIδ−/− myocytes developed significantly less H<span class="hlt">2</span>O<span class="hlt">2</span>-induced arrhythmias, and were more resistant to hypercontracture. Opposite results (increased late INa, [<span class="hlt">Na</span>]i and [<span class="hlt">Ca]i</span> accumulation) were obtained by overexpression of CaMKIIδ in rabbit myocytes (adenoviral gene transfer) reversible with CaMKII inhibition (10 µmol/L KN93 or 0.1 µmol/L AIP). Conclusion Free [<span class="hlt">Ca]i</span> and a functional SR are required for ROS activation of CaMKII. ROS-activated CaMKIIδ enhances late INa, which may lead to cellular <span class="hlt">Na</span> and Ca overload. This may be of relevance in HF, where enhanced ROS production meets increased CaMKII expression. PMID:21252154</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFM.A23A0181H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFM.A23A0181H"><span>Discussion of vicarious calibration of GOSAT/TANSO-<span class="hlt">CAI</span> UV-band (380nm) and aerosol retrieval in wildfire region in the OCO-<span class="hlt">2</span> and GOSAT observation campaign at Railroad Valley in 2016</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hashimoto, M.; Kuze, A.; Bruegge, C. J.; Shiomi, K.; Kataoka, F.; Kikuchi, N.; Arai, T.; Kasai, K.; Nakajima, T.</p> <p>2016-12-01</p> <p>The GOSAT (Greenhouse Gases Observing Satellite) / TANSO-<span class="hlt">CAI</span> (Cloud and Aerosol Imager, <span class="hlt">CAI</span>) is an imaging sensor to measure cloud and aerosol properties and observes reflected sunlight from the atmosphere and surface of the ground. The sensor has four bands from near ultraviolet (near-UV) to shortwave infrared, 380, 674, 870 and 1600nm. The field of view size is 0.5 km for band-1 through band-3, and 1.5km for band-4. Band-1 (380nm) is one of unique function of the <span class="hlt">CAI</span>. The near-UV observation offers several advantages for the remote sensing of aerosols over land: Low reflectance of most surfaces; Sensitivity to absorbing aerosols; Absorption of trace gases is weak (Höller et al., 2004). <span class="hlt">CAI</span> UV-band is useful to distinguish absorbing aerosol (smoke) from cloud. GOSAT-<span class="hlt">2</span>/TANSO-<span class="hlt">CAI</span>-<span class="hlt">2</span> that will be launched in the future also has UV-bands, 340 and 380nm. We carried out an experiment to calibrate <span class="hlt">CAI</span> UV-band radiance using data taken in a field campaign of OCO-<span class="hlt">2</span> and GOSAT at Railroad Valley in 2016. The campaign period is June 27 to July 3 in 2016. We measured surface reflectance by using USB4000 Spectrometer with 74-UV collimating lens (Ocean Optics) and Spectralon (Labsphere). USB4000 is a UV spectrometer, and its measurement range from 300 to 520nm. We simulated <span class="hlt">CAI</span> UV-band radiance using a vector type of radiation transfer code, i.e. including polarization calculation, pstar3 (Ota et al., 2010) using measured surface reflectance and atmospheric data, pressure and relative humidity by radiosonde in the same campaign, and aerosol optical depth by AERONET, etc. Then, we evaluated measured UV radiances with the simulated data. We show the result of vicarious calibration of <span class="hlt">CAI</span> UV-band in the campaign, and discuss about this method for future sensor, <span class="hlt">CAI</span>-<span class="hlt">2</span>. Around the campaign period, there was wildfire around Los Angeles, and aerosol optical thickness (AOT) observed by AERONET at Rail Road valley and Caltech sites is also high. We tried to detect and retrieve aerosol</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://files.eric.ed.gov/fulltext/ED097910.pdf','ERIC'); return false;" href="http://files.eric.ed.gov/fulltext/ED097910.pdf"><span><span class="hlt">CAI</span>: Overcoming Attitude Barriers.</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>Netusil, Anton J.; Kockler, Lois H.</p> <p></p> <p>During each of two school quarters, approximately 60 college students enrolled in a mathematics course were randomly assigned to an experimental group or a control group. The control group received instruction by the lecture method only; the experimental group received the same instruction, except that six computer-assisted instruction (<span class="hlt">CAI</span>) units…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://eric.ed.gov/?q=future+AND+AI&pg=7&id=ED322479','ERIC'); return false;" href="https://eric.ed.gov/?q=future+AND+AI&pg=7&id=ED322479"><span>The Relative Effectiveness of Computer Assisted Instruction (<span class="hlt">CAI</span>) for Teaching Students To Read English.</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>Thompson, Richard A.</p> <p></p> <p>In a review of research on computer assisted instruction (<span class="hlt">CAI</span>) related to reading, evidence collected provides tentative conclusions about <span class="hlt">CAI</span> effectiveness. <span class="hlt">CAI</span> was effective as an instructional medium in the surveyed studies. In a number of instances, <span class="hlt">CAI</span> groups achieved higher scores than the control groups. Some studies indicated that CAI…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19950012899&hterms=Mg+Ca&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3DMg%2BCa','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19950012899&hterms=Mg+Ca&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3DMg%2BCa"><span>Chronology of chrondrule and <span class="hlt">CAI</span> formation: Mg-Al isotopic evidence</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Macpherson, G. J.; Davis, A. M.</p> <p>1994-01-01</p> <p>Details of the chondrule and Ca-Al-rich inclusion (<span class="hlt">CAI</span>) formation during the earliest history of the solar system are imperfectly known. Because <span class="hlt">CAI</span>'s are more 'refractory' than ferromagnesian chondrules and have the lowest recorded initial Sr-87/Sr-86 ratios of any solar system materials, the expectation is that <span class="hlt">CAI</span>'s formed earlier than chondrules. But it is not known, for example, if <span class="hlt">CAI</span> formation had stopped by the time chondrule formation began. Conventional (absolute) age-dating techniques cannot adequately resolve small age differences (less than 10(exp 6) years) between objects of such antiquity. One approach has been to look at systematic differences in the daughter products of short-lived radionuclides such as Al-26 and I-129. Unfortunately, neither system appears to be 'well-behaved.' One possible reason for this circumstance is that later secondary events have partially reset the isotopic systems, but a viable alternative continues to be large-scale (nebular) heterogeneity in initial isotopic abundances, which would of course render the systems nearly useless as chronometers. In the past two years the nature of this problem has been redefined somewhat. Examination of the Al-Mg isotopic database for all <span class="hlt">CAI</span>'s suggests that the vast majority of inclusions originally had the same initial Al-26/Al-27 abundance ratio, and that the ill-behaved isotopic systematics now observed are the results of later partial reequilibration due to thermal processing. Isotopic heterogeneities did exist in the nebula, as demonstrated by the existence of so-called FUN inclusions in CV3 chondrites and isotopically anomalous hibonite grains in CM<span class="hlt">2</span> chondrites, which had little or no live Al-26 at the time of their formation. But, among the population of CV3 inclusions at least, FUN inclusions appear to have been a relatively minor nebular component.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://files.eric.ed.gov/fulltext/ED136826.pdf','ERIC'); return false;" href="http://files.eric.ed.gov/fulltext/ED136826.pdf"><span>An Intelligent <span class="hlt">CAI</span> Monitor and Generative Tutor. Final Report.</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>Koffman, Elliot B.; Perry, James</p> <p></p> <p>This final report summarizes research findings and presents a model for generative computer assisted instruction (<span class="hlt">CAI</span>) with respect to its usefulness in the classroom environment. Methods used to individualize instruction, and the evolution of a procedure used to select a concept for presentation to a student with the generative <span class="hlt">CAI</span> system are…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20170005741','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20170005741"><span>NWA10758: A New CV3 Chondrite Bearing a Giant <span class="hlt">CAI</span> with Hibonite-Rich Wark-Lovering Rim</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Ross, D. K.; Simon, J. I.; Zolensky, M.</p> <p>2017-01-01</p> <p>Northwest Africa (NWA) 10758 is a newly identified carbonaceous chondrite that is a Bali-like oxidized CV3. The large Ca-Al rich inclusion (<span class="hlt">CAI</span>) in this sample is approx. <span class="hlt">2</span>.4 x 1.4 cm. The <span class="hlt">CAI</span> is transitional in composition between type A and type B, with interior mineralogy dominated by melilite, plus less abundant spinel and Al-Ti rich diopside, and only very minor anorthite (Fig. 1A). This <span class="hlt">CAI</span> is largely free of secondary alteration in the exposed section we examined, with almost no nepheline, sodalite or Ca-Fe silicates. The Wark-Lovering (WL) rim on this <span class="hlt">CAI</span> is dominated by hibonite, with lower abundances of spinel and perovskite, and with hibonite locally overlain by melilite plus perovskite (as in Fig. 1B). Note that the example shown in 1B is exceptional. Around most of the <span class="hlt">CAI</span>, hibonite + spinel + perovskite form the WL rim, without overlying melilite. The WL rim can be unusually thick, ranging from approx. 20 microns up to approx. 150 microns. A well-developed, stratified accretionary rim infills embayments of the <span class="hlt">CAI</span>, and thins over protuberances in the convoluted <span class="hlt">CAI</span> surface.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://files.eric.ed.gov/fulltext/EJ1097442.pdf','ERIC'); return false;" href="http://files.eric.ed.gov/fulltext/EJ1097442.pdf"><span>Effectiveness of <span class="hlt">CAI</span> Package on Achievement in Physics of IX Standard Students</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>Maheswari, I. Uma; Ramakrishnan, N.</p> <p>2015-01-01</p> <p>The present study is an experimental one in nature, to find out the effectiveness of <span class="hlt">CAI</span> package on in Physics of IX std. students. For this purpose a <span class="hlt">CAI</span> package was developed and validated. The validated <span class="hlt">CAI</span> package formed an independent variable of this study. The dependent variable is students' achievements in physics content. In order to find…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/16621162','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/16621162"><span>Single and combined effects of carbamazepine and vinpocetine on depolarization-induced changes in <span class="hlt">Na</span>+, Ca<span class="hlt">2</span>+ and glutamate release in hippocampal isolated nerve endings.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Sitges, María; Chiu, Luz María; Nekrassov, Vladimir</p> <p>2006-07-01</p> <p>The single and combined effects of carbamazepine and vinpocetine on the release of the excitatory amino acid neurotransmitter glutamate, on the rise in internal <span class="hlt">Na</span>+ (<span class="hlt">Na</span>(i), as determined with SBFI), and on the rise in internal Ca<span class="hlt">2</span>+ (<span class="hlt">Ca(i</span>), as determined with fura-<span class="hlt">2</span>) induced by an increased permeability of presynaptic <span class="hlt">Na</span>+ channels, with veratridine, or by an increased permeability of presynaptic Ca<span class="hlt">2</span>+ channels with high K+, were investigated in isolated hippocampal nerve endings. The present study shows that carbamazepine and vinpocetine, both inhibit dose dependently the release of preloaded [3H]Glu induced by veratridine. However, carbamazepine is two orders of magnitude less potent than vinpocetine. The calculated IC(50)'s for carbamazepine and vinpocetine to inhibit veratridine-induced [3H]Glu release are 200 and <span class="hlt">2</span> microM, respectively. Consistently 150 microM carbamazepine and 1.5 microM vinpocetine reduce the veratridine-induced rise in <span class="hlt">Na</span>(i) in a similar extent. The single effects of carbamazepine and of vinpocetine on the presynaptic <span class="hlt">Na</span>+ channel mediated responses, namely the rise in <span class="hlt">Na</span>(i) and the release of Glu induced by veratridine, are additive. Responses that depend on the entrance of external Ca<span class="hlt">2</span>+ via presynaptic Ca<span class="hlt">2</span>+ channels, such as the release of [3H]Glu and the rise in <span class="hlt">Ca(i</span>) induced by high K+, are insensitive to 300 microM carbamazepine and slightly reduced by 5 microM vinpocetine. It is concluded that the additive effects of carbamazepine, which is one of the most common antiepileptic drugs, and vinpocetine that besides its known neuroprotective action and antiepileptic potential is a memory enhancer, may perhaps be advantageous in the treatment of epileptic patients.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20170006931','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20170006931"><span>Northwest Africa 10758: A New CV3 Chondrite Bearing a Giant <span class="hlt">CAI</span> with Hibonite-Rich Wark-Lovering Rim</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Ross, D. K.; Simon, J. I.; Zolensky, M.</p> <p>2017-01-01</p> <p>Northwest Africa (NWA) 10758 is a newly identified carbonaceous chondrite that is a Bali-like oxidized CV3. The large Ca-Al rich inclusion (<span class="hlt">CAI</span>) in this sample is approx. <span class="hlt">2</span>.4 x 1.4 cm. The <span class="hlt">CAI</span> is transitional in composition between type A and type B, with interior mineralogy dominated by melilite, plus less abundant spinel and Al-Ti rich diopside, and only very minor anorthite (Fig. 1A). This <span class="hlt">CAI</span> is largely free of secondary alteration in the exposed section we examined, with almost no nepheline, sodalite or Ca-Fe silicates. The Wark-Lovering (WL) rim on this <span class="hlt">CAI</span> is dominated by hibonite, with lower abundances of spinel and perovskite, and with hibonite locally overlain by melilite plus perovskite (as in Fig. 1B). Note that the example shown in 1B is exceptional. Around most of the <span class="hlt">CAI</span>, hibonite + spinel + perovskite form the WL rim, without overlying melilite. The WL rim can be unusually thick, ranging from approx.20 microns up to approx. 150 microns. A well-developed, stratified accretionary rim infills embayments of the <span class="hlt">CAI</span>, and thins over protuberances in the convoluted <span class="hlt">CAI</span> surface.</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('https://www.ncbi.nlm.nih.gov/pubmed/15235824','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/15235824"><span>A phase I trial of pharmacokinetic modulation of carboxyamidotriazole (<span class="hlt">CAI</span>) with ketoconazole in patients with advanced cancer.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Desai, Apurva A; Innocenti, Federico; Janisch, Linda; DeMario, Mark; Shepard, Dale; Ramirez, Jacqueline; Fleming, Gini F; Ratain, Mark J</p> <p>2004-11-01</p> <p>Carboxyamidotriazole (<span class="hlt">CAI</span>) is a novel antineoplastic agent in clinical development with limited oral bioavailability. In vitro, ketoconazole has been demonstrated to inhibit CYP3A4-mediated metabolism of <span class="hlt">CAI</span>. We performed this phase I trial to determine if ketoconazole-mediated CYP3A4 inhibition would lead to favorable alteration of <span class="hlt">CAI</span> pharmacokinetics, and to evaluate the safety, toxicity and tolerability of the proposed combination. Forty-seven patients were treated using a standard three patients per cohort <span class="hlt">CAI</span> dose-escalation scheme. In cycle 1, <span class="hlt">CAI</span> was administered alone on day-6 followed by a single dose of ketoconazole (200 mg) on day 0. <span class="hlt">CAI</span> and ketoconazole (200 mg/day) were subsequently coadministered on days 1 and 3-28. Plasma samples for pharmacokinetic analysis were obtained following the doses on days-6 and 1. All subsequent cycles were of 28-day duration, and consisted of daily <span class="hlt">CAI</span> and ketoconazole coadministration. Pharmacokinetic analysis was performed on samples from 44 patients. In most patients administration of ketoconazole produced an increase in <span class="hlt">CAI</span> AUC and Cmax with a decrease in <span class="hlt">CAI</span> clearance. Seven patients experienced stable disease for up to 12 months. Gastrointestinal and constitutional toxicities were the most common toxicities. Coadministration of <span class="hlt">CAI</span> with ketoconazole increased <span class="hlt">CAI</span> exposure in most of the patients without altering the toxicity profile of <span class="hlt">CAI</span>. The highest <span class="hlt">CAI</span> dose administered on the trial was 300 mg/day. The clinical utility of such a modulation strategy might be explored in future clinical trials of <span class="hlt">CAI</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/15011596','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/15011596"><span>Experimental Determination of Li, Be and B Partitioning During <span class="hlt">CAI</span> Crystallization</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>Ryerson, F J; Brenan, J M; Phinney, D L</p> <p>2005-01-12</p> <p>The main focus of the work is to develop a better understanding of the distribution of the elements B, Be and Li in melilite, fassaitic clinop clinopy-roxene, anorthite and spinel, which are the primary constituents of calcium-aluminum-rich inclusions (<span class="hlt">CAIs</span>). These elements are the parent or decay products of short-lived nuclides (specifically, {sup 7}Be and {sup 10}Be) formed by cosmic ray spallation reactions on silicon and oxygen. Recent observations suggest that some <span class="hlt">CAIs</span> contain ''fossil'' {sup 7}Be and {sup 10}Be in the form of ''excess'' amounts of their decay products (B and Li). The exact timing of {sup 7}Be and {supmore » 10}Be production is unknown, but if it occurred early in <span class="hlt">CAI</span> history, it could constrain the birthplace of <span class="hlt">CAIs</span> to be within a limited region near the infant sun. Other interpretations are possible, however, and bear little significance to early <span class="hlt">CAI</span> genesis. In order to interpret the anomalies as being ''primary'', and thus originating at high temperature, information on the intermineral partitioning of both parent and daughter elements is required.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://files.eric.ed.gov/fulltext/EJ1068387.pdf','ERIC'); return false;" href="http://files.eric.ed.gov/fulltext/EJ1068387.pdf"><span>Effect of <span class="hlt">CAI</span> on Achievement of LD Students in English</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>Sivaram, R. T.; Ramar, R.</p> <p>2014-01-01</p> <p>The present experimental study was undertaken with three objectives in view, (i) to identify students with language learning disabilities (ii) to develop <span class="hlt">CAI</span> software to teach LD students through computer-assisted instruction and (iii) to measure the effectiveness of <span class="hlt">CAI</span> with special reference to LD students. Two matched groups of LD students were…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20170005632','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20170005632"><span>Multiple Nebular Gas Reservoirs Recorded by Oxygen Isotope Variation in a Spinel-rich <span class="hlt">CAI</span> in CO3 MIL 090019</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Simon, J. I.; Simon, S. B.; Nguyen, A. N.; Ross, D. K.; Messenger, S.</p> <p>2017-01-01</p> <p>We conducted NanoSIMS O-isotopic imaging of a primitive spinel-rich <span class="hlt">CAI</span> spherule (27-<span class="hlt">2</span>) from the MIL 090019 CO3 chondrite. Inclusions such as 27-<span class="hlt">2</span> are proposed to record inner nebula processes during an epoch of rapid solar nebula evolution. Mineralogical and textural analyses suggest that this <span class="hlt">CAI</span> formed by high temperature reactions, partial melting, and condensation. This <span class="hlt">CAI</span> exhibits radial O-isotopic heterogeneity among multiple occurrences of the same mineral, reflecting interactions with distinct nebular O-isotopic reservoirs.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://files.eric.ed.gov/fulltext/ED428702.pdf','ERIC'); return false;" href="http://files.eric.ed.gov/fulltext/ED428702.pdf"><span>An Object-Oriented Architecture for a Web-Based <span class="hlt">CAI</span> System.</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>Nakabayashi, Kiyoshi; Hoshide, Takahide; Seshimo, Hitoshi; Fukuhara, Yoshimi</p> <p></p> <p>This paper describes the design and implementation of an object-oriented World Wide Web-based <span class="hlt">CAI</span> (Computer-Assisted Instruction) system. The goal of the design is to provide a flexible <span class="hlt">CAI</span>/ITS (Intelligent Tutoring System) framework with full extendibility and reusability, as well as to exploit Web-based software technologies such as JAVA, ASP (a…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://eric.ed.gov/?q=Fortran+AND+programming&pg=4&id=ED064917','ERIC'); return false;" href="https://eric.ed.gov/?q=Fortran+AND+programming&pg=4&id=ED064917"><span><span class="hlt">CAI</span>-BASIC: A Program to Teach the Programming Language BASIC.</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>Barry, Thomas Anthony</p> <p></p> <p>A computer-assisted instruction (<span class="hlt">CAI</span>) program was designed which fulfills the objectives of teaching a simple programing language, interpreting student responses, and executing and editing student programs. The <span class="hlt">CAI</span>-BASIC program is written in FORTRAN IV and executes on IBM-2741 terminals while running under a time-sharing system on an IBM-360-70…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/21385711','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21385711"><span>Oxygen isotope variations at the margin of a <span class="hlt">CAI</span> records circulation within the solar nebula.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Simon, Justin I; Hutcheon, Ian D; Simon, Steven B; Matzel, Jennifer E P; Ramon, Erick C; Weber, Peter K; Grossman, Lawrence; DePaolo, Donald J</p> <p>2011-03-04</p> <p>Micrometer-scale analyses of a calcium-, aluminum-rich inclusion (<span class="hlt">CAI</span>) and the characteristic mineral bands mantling the <span class="hlt">CAI</span> reveal that the outer parts of this primitive object have a large range of oxygen isotope compositions. The variations are systematic; the relative abundance of (16)O first decreases toward the <span class="hlt">CAI</span> margin, approaching a planetary-like isotopic composition, then shifts to extremely (16)O-rich compositions through the surrounding rim. The variability implies that <span class="hlt">CAIs</span> probably formed from several oxygen reservoirs. The observations support early and short-lived fluctuations of the environment in which <span class="hlt">CAIs</span> formed, either because of transport of the <span class="hlt">CAIs</span> themselves to distinct regions of the solar nebula or because of varying gas composition near the proto-Sun.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=107209','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=107209"><span>Regulation of the Carnitine Pathway in Escherichia coli: Investigation of the <span class="hlt">cai</span>-fix Divergent Promoter Region</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Buchet, Anne; Eichler, Knut; Mandrand-Berthelot, Marie-Andrée</p> <p>1998-01-01</p> <p>The divergent structural operons <span class="hlt">cai</span>TABCDE and fixABCX of Escherichia coli are required for anaerobic carnitine metabolism. Transcriptional monocopy lacZ fusion studies showed that both operons are coexpressed during anaerobic growth in the presence of carnitine, respond to common environmental stimuli (like glucose and nitrate), and are modulated positively by the same general regulators, CRP and FNR, and negatively by H-NS. Overproduction of the <span class="hlt">Cai</span>F specific regulatory protein mediating the carnitine signal restored induction in an fnr mutant, corresponding to its role as the primary target for anaerobiosis. Transcript analysis identified two divergent transcription start points initiating 289 bp apart. DNase I footprinting revealed three sites with various affinities for the binding of the cAMP-CRP complex inside this regulatory region. Site-directed mutagenesis experiments indicated that previously reported perfect CRP motif 1, centered at −41.5 of the <span class="hlt">cai</span> transcriptional start site, plays a direct role in the sole <span class="hlt">cai</span> activation. In contrast, mutation in CRP site <span class="hlt">2</span>, positioned at −69.5 of the fix promoter, caused only a threefold reduction in fix expression. Thus, the role of the third CRP site, located at −126.5 of fix, might be to reinforce the action of site <span class="hlt">2</span>. A critical 50-bp cis-acting sequence overlapping the fix mRNA start site was found, by deletion analysis, to be necessary for <span class="hlt">cai</span> transcription. This region is thought to be involved in transduction of the signal mediated by the <span class="hlt">Cai</span>F regulator. PMID:9573142</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19950042229&hterms=FeTiO3&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3DFeTiO3','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19950042229&hterms=FeTiO3&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3DFeTiO3"><span>An ion microprobe study of <span class="hlt">CAIs</span> from CO3 meteorites. [Abstract only</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Russell, S. S.; Greenwood, R. C.; Fahey, A. J.; Huss, G. R.; Wasserburg, G. J.</p> <p>1994-01-01</p> <p>When attempting to interpret the history of Ca, Al-rich inclusions (<span class="hlt">CAIs</span>) it is often difficult to distinguish between primary features inherited from the nebula and those produced during secondary processing on the parent body. We have undertaken a systematic study of <span class="hlt">CAIs</span> from 10 CO chondrites, believed to represent a metamorphic sequence with the goal of distinguishing primary and secondary features. ALHA 77307 (3.0), Colony (3.0), Kainsaz (3.1), Felix (3.<span class="hlt">2</span>), ALH 82101 (3.3), Ornans (3.3), Lance (3.4), ALHA 77003 (3.5), Warrenton (3.6), and Isna (3.7) were examined by Scanning Electron Microscopy (SEM) and optical microscopy. We have identified 141 <span class="hlt">CAIs</span> within these samples, and studied in detail the petrology of 34 inclusions. The primary phases in the lower petrologic types are spinel, melilite, and hibonite. Perovskite, FeS, ilmenite, anorthite, kirschsteinite, and metallic Fe are present as minor phases. Melilite becomes less abundant in higher petrologic types and was not detected in chondrites of type 3.5 and above, confirming previous reports that this mineral easily breaks down during heating. Iron, an element that would not be expected to condense at high temperatures, has a lower abundance in spinel from low-petrologic-type meteorites than those of higher grade, and CaTiO3 is replaced by FeTiO3 in meteorites of higher petrologic type. The abundance of <span class="hlt">CAIs</span> is similar in each meteorite. Eight inclusions have been analyzed by ion probe. The results are summarized. The results obtained to date show that <span class="hlt">CAIs</span> in CO meteorites, like those from other meteorite classes, contain Mg* and that Mg in some inclusions has been redistributed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22078296-variable-extreme-irradiation-conditions-early-solar-system-inferred-from-initial-abundance-sup-isheyevo-cais','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22078296-variable-extreme-irradiation-conditions-early-solar-system-inferred-from-initial-abundance-sup-isheyevo-cais"><span>VARIABLE AND EXTREME IRRADIATION CONDITIONS IN THE EARLY SOLAR SYSTEM INFERRED FROM THE INITIAL ABUNDANCE OF {sup 10}Be IN ISHEYEVO <span class="hlt">CAIs</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>Gounelle, Matthieu; Chaussidon, Marc; Rollion-Bard, Claire, E-mail: gounelle@mnhn.fr</p> <p>2013-02-01</p> <p>A search for short-lived {sup 10}Be in 21 calcium-aluminum-rich inclusions (<span class="hlt">CAIs</span>) from Isheyevo, a rare CB/CH chondrite, showed that only 5 <span class="hlt">CAIs</span> had {sup 10}B/{sup 11}B ratios higher than chondritic correlating with the elemental ratio {sup 9}Be/{sup 11}B, suggestive of in situ decay of this key short-lived radionuclide. The initial ({sup 10}Be/{sup 9}Be){sub 0} ratios vary between {approx}10{sup -3} and {approx}10{sup -<span class="hlt">2</span>} for <span class="hlt">CAI</span> 411. The initial ratio of <span class="hlt">CAI</span> 411 is one order of magnitude higher than the highest ratio found in CV3 <span class="hlt">CAIs</span>, suggesting that the more likely origin of <span class="hlt">CAI</span> 411 {sup 10}Be is early solar systemmore » irradiation. The low ({sup 26}Al/{sup 27}Al){sub 0} [{<=} 8.9 Multiplication-Sign 10{sup -7}] with which <span class="hlt">CAI</span> 411 formed indicates that it was exposed to gradual flares with a proton fluence of a few 10{sup 19} protons cm{sup -<span class="hlt">2</span>}, during the earliest phases of the solar system, possibly the infrared class 0. The irradiation conditions for other <span class="hlt">CAIs</span> are less well constrained, with calculated fluences ranging between a few 10{sup 19} and 10{sup 20} protons cm{sup -<span class="hlt">2</span>}. The variable and extreme value of the initial {sup 10}Be/{sup 9}Be ratios in carbonaceous chondrite <span class="hlt">CAIs</span> is the reflection of the variable and extreme magnetic activity in young stars observed in the X-ray domain.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017GeCoA.201...65M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017GeCoA.201...65M"><span>High precision Al-Mg systematics of forsterite-bearing Type B <span class="hlt">CAIs</span> from CV3 chondrites</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>MacPherson, G. J.; Bullock, E. S.; Tenner, T. J.; Nakashima, D.; Kita, N. T.; Ivanova, M. A.; Krot, A. N.; Petaev, M. I.; Jacobsen, S. B.</p> <p>2017-03-01</p> <p>In order to further elucidate possible temporal relationships between different varieties of calcium-, aluminum-rich inclusions (<span class="hlt">CAIs</span>), we measured the aluminum-magnesium isotopic systematics of seven examples of the rare type known as forsterite-bearing Type B (FoB) inclusions from four different CV3 carbonaceous chondrites: Allende, Efremovka, NWA 3118, and Vigarano. The primary phases (forsterite, Al-Ti-rich diopside, spinel, melilite, and anorthite) in each inclusion were analyzed in situ using high-precision secondary ion mass-spectrometry (SIMS). In all cases, minerals with low Al/Mg ratios (all except anorthite) yield well-defined internal Al-Mg isochrons, with a range of initial 26Al/27Al ratios [(26Al/27Al)0] ranging from (5.30 ± 0.22) × 10-5 down to (4.17 ± 0.43) × 10-5. Anorthite in all cases is significantly disturbed relative to the isochrons defined by the other phases in the same <span class="hlt">CAIs</span>, and in several cases contains no resolved excesses of radiogenic 26Mg (δ26Mg∗) even at 27Al/24Mg ratios greater than 1000. The fact that some FoBs preserve (26Al/27Al)0 of ∼5.<span class="hlt">2</span> × 10-5, close to the canonical value of (5.23 ± 0.13) × 10-5 inferred from bulk magnesium-isotope measurements of CV <span class="hlt">CAIs</span> (B. Jacobsen et al., 2008), demonstrates that FoBs began forming very early, contemporaneous with other more-refractory <span class="hlt">CAIs</span>. The range of (26Al/27Al)0 values further shows that FoBs continued to be reprocessed over ∼200,000 years of nebular history, consistent with results obtained for other types of igneous <span class="hlt">CAIs</span> in CV chondrites. The absence of any correlation between of <span class="hlt">CAI</span> + FoB formation or reprocessing times with bulk composition or <span class="hlt">CAI</span> type means that there is no temporal evolutionary sequence between the diverse <span class="hlt">CAI</span> types. The initial δ26Mg∗ value in the most primitive FoB (SJ101) is significantly lower than the canonical solar system value of -0.040 ± 0.029‰.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20150010430','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20150010430"><span>Microstructural Investigation of a Wark-Lovering Rim on a Vigarano <span class="hlt">CAI</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Han, J.; Keller, L. P.; Needham, A. W.; Messenger, S.; Simon, J. I.</p> <p>2015-01-01</p> <p>Wark-Lovering (WL) rims are thin multi-layered mineral sequences that surround many <span class="hlt">CAIs</span>. These rim layers consist of the primary minerals found in the <span class="hlt">CAI</span> interiors, but vary in their mineralogy. Several models for their origin have been proposed including condensation, reaction with a nebular gas, evaporation, or combinations of these. However, there still is little consensus on how and when the rims formed. Here, we describe the microstructure and mineralogy of a WL rim on a type B <span class="hlt">CAI</span> from the Vigarano CV(sub red) chondrite using FIB/TEM to better understand the astrophysical significance of WL rim formation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://eric.ed.gov/?q=Runes&id=ED218939','ERIC'); return false;" href="https://eric.ed.gov/?q=Runes&id=ED218939"><span><span class="hlt">CAI</span> in Advanced Literature Class.</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>Hinton, Norman</p> <p>1981-01-01</p> <p>Ways that computer assisted instruction (<span class="hlt">CAI</span>) can be useful in teaching English at upperclass and graduate levels are considered, with illustrations from PLATO lessons that have been composed and programmed. One lesson takes advantage of PLATO's graphic design capabilities, which enabled the teacher to design the runic figures and to show them in…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20090020501','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20090020501"><span>Rare Earth Element Measurements of Melilite and Fassaite in Allende <span class="hlt">Cai</span> by Nanosims</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Ito, M.; Messenger, Scott</p> <p>2009-01-01</p> <p>The rare earth elements (REEs) are concentrated in <span class="hlt">CAIs</span> by approx. 20 times the chondritic average [e.g., 1]. The REEs in <span class="hlt">CAIs</span> are important to understand processes of <span class="hlt">CAI</span> formation including the role of volatilization, condensation, and fractional crystallization [1,<span class="hlt">2</span>]. REE measurements are a well established application of ion microprobes [e.g., 3]. However the spatial resolution of REE measurements by ion microprobe (approx.20 m) is not adequate to resolve heterogeneous distributions of REEs among/within minerals. We have developed methods for measuring REE with the NanoSIMS 50L at smaller spatial scales. Here we present our initial measurements of REEs in melilite and fassaite in an Allende Type-A <span class="hlt">CAI</span> with the JSC NanoSIMS 50L. We found that the key parameters for accurate REE abundance measurements differ between the NanoSIMS and conventional SIMS, in particular the oxide-to-element ratios, the relative sensitivity factors, the energy distributions, and requisite energy offset. Our REE abundance measurements of the 100 ppm REE diopside glass standards yielded good reproducibility and accuracy, 0.5-<span class="hlt">2</span>.5 % and 5-25 %, respectively. We determined abundances and spatial distributions of REEs in core and rim within single crystals of fassaite, and adjacent melilite with 5-10 m spatial resolution. The REE abundances in fassaite core and rim are 20-100 times CI abundance but show a large negative Eu anomaly, exhibiting a well-defined Group III pattern. This is consistent with previous work [4]. On the other hand, adjacent melilite shows modified Group II pattern with no strong depletions of Eu and Yb, and no Tm positive anomaly. REE abundances (<span class="hlt">2</span>-10 x CI) were lower than that of fassaite. These patterns suggest that fassaite crystallized first followed by a crystallization of melilite from the residual melt. In future work, we will carry out a correlated study of O and Mg isotopes and REEs of the <span class="hlt">CAI</span> in order to better understand the nature and timescales of its</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20170006939','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20170006939"><span>Oxygen, Magnesium, and Aluminum Isotopes in the Ivuna <span class="hlt">CAI</span>: Re-Examining High-Temperature Fractionations in CI Chondrites</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Frank, D. R.; Huss, G. R.; Nagashima, K.; Zolensky, M. E.; Le, L.</p> <p>2017-01-01</p> <p>CI chondrites are thought to approximate the bulk solar system composition since they closely match the composition of the solar photosphere. Thus, chemical differences between a planetary object and the CI composition are interpreted to result from fractionations of a CI starting composition. This interpretation is often made despite the secondary mineralogy of CI chondrites, which resulted from low-T aqueous alteration on the parent asteroid(s). Prevalent alteration and the relatively large uncertainties in the photospheric abundances (approx. +/-5-10%) permit chemical fractionation of CI chondrites from the bulk solar system, if primary chondrules and/or <span class="hlt">CAIs</span> have been altered beyond recognition. Isolated olivine and pyroxene grains that range from approx. 5 microns to several hundred microns have been reported in CI chondrites, and acid residues of Orgueil were found to contain refractory oxides with oxygen isotopic compositions matching <span class="hlt">CAIs</span>. However, the only <span class="hlt">CAI</span> found to be unambiguously preserved in a CI chondrite was identified in Ivuna. The Ivuna <span class="hlt">CAI</span>'s primary mineralogy, small size (approx.170 microns), and fine-grained igneous texture classify it as a compact type A. Aqueous alteration infiltrated large portions of the <span class="hlt">CAI</span>, but other regions remain pristine. The major primary phases are melilite (Ak 14-36 ), grossmanite (up to 20.8 wt.% TiO <span class="hlt">2</span> ), and spinel. Both melilite and grossmanite have igneous textures and zoning patterns. An accretionary rim consists primarily of olivine (Fa <span class="hlt">2</span>-17 ) and low-Ca pyroxene (Fs <span class="hlt">2</span>-10 ), which could be either surviving CI<span class="hlt">2</span> material or a third lithology.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2004AGUFM.P33A1006Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2004AGUFM.P33A1006Y"><span>Unraveling the Environmental Record of the Early Solar System: High Precision Laser Ablation Al-Mg Isotopes of Igneous <span class="hlt">CAIs</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Young, E. D.; Simon, J. I.; Russell, S. S.; Tonui, E.; Krot, A.</p> <p>2004-12-01</p> <p>Variations in intrinsic Mg isotope compositions provide a potentially rich record of the physiochemical evolution of <span class="hlt">CAIs</span>. Moreover, Mg excesses from the short-lived 26Al chronometer can be used to constrain when these processes occurred; e.g., during the nebular phase and/or during the development of planetisimals (< 4 Myr). We obtained in situ UV (213 nm) laser ablation MC-ICPMS measurements of Al and Mg isotope ratios within core-to-rim traverses of igneous <span class="hlt">CAIs</span> to place temporal constraints on when features of <span class="hlt">CAIs</span> formed. Results provide tests of models for the chemical and isotopic evolution of <span class="hlt">CAIs</span> involving volatilization and recondensation of elements in the solar nebula. We studied five CV3 <span class="hlt">CAIs</span>, including Allende 3576-1 "b", Allende M5, Leoville 144A, Leoville MRS3, and Efremovka E44. Our sample-standard comparison approach affords a precision <0.<span class="hlt">2</span> \\permil per amu (<span class="hlt">2</span>s) for intrinsic Mg isotope measurements and <0.3 \\permil (<span class="hlt">2</span>s) for measured 26Mg excesses. Intra-object variation in \\delta25Mg exists with values ranging from as low as -<span class="hlt">2</span> \\permil and as high as +8 \\permil (compared to DSM3). The distinct Mg isotope patterns in the <span class="hlt">CAIs</span> are difficult to explain by a single process or within a single nebular environment and likely require changing conditions or transfer of <span class="hlt">CAIs</span> from one nebular environment to another. The ˜pristine Mg isotope profile of Leoville 144A is compared to results produced by implicit finite difference modeling. Model curves reflect isotopic fractionation at the moving surface of a shrinking molten sphere coupled with diffusion-limited transport within the sphere. We find that using mass-dependant diffusivities increases \\delta25Mg with evaporation, but does not produce the tight curvature in the edgeward increases in \\delta25Mg characteristic of Leoville 144A. Three <span class="hlt">CAIs</span> that exhibit edgeward \\delta25Mg decreases are well described by diffusion in a Mg-rich chondritic environment suggestive of nebular temperatures and</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26852403','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26852403"><span>Cognitive Assessment Interview (<span class="hlt">CAI</span>): Validity as a co-primary measure of cognition across phases of schizophrenia.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Ventura, Joseph; Subotnik, Kenneth L; Ered, Arielle; Hellemann, Gerhard S; Nuechterlein, Keith H</p> <p>2016-04-01</p> <p>Progress has been made in developing interview-based measures for the assessment of cognitive functioning, such as the Cognitive Assessment Interview (<span class="hlt">CAI</span>), as co-primary measures that compliment objective neurocognitive assessments and daily functioning. However, a few questions remain, including whether the relationships with objective cognitive measures and daily functioning are high enough to justify the <span class="hlt">CAI</span> as an co-primary measure and whether patient-only assessments are valid. Participants were first-episode schizophrenia patients (n=60) and demographically-similar healthy controls (n=35), chronic schizophrenia patients (n=38) and demographically similar healthy controls (n=19). Participants were assessed at baseline with an interview-based measure of cognitive functioning (<span class="hlt">CAI</span>), a test of objective cognitive functioning, functional capacity, and role functioning at baseline, and in the first episode patients again 6 months later (n=28). <span class="hlt">CAI</span> ratings were correlated with objective cognitive functioning, functional capacity, and functional outcomes in first-episode schizophrenia patients at similar magnitudes as in chronic patients. Comparisons of first-episode and chronic patients with healthy controls indicated that the <span class="hlt">CAI</span> sensitively detected deficits in schizophrenia. The relationship of <span class="hlt">CAI</span> Patient-Only ratings with objective cognitive functioning, functional capacity, and daily functioning were comparable to <span class="hlt">CAI</span> Rater scores that included informant information. These results confirm in an independent sample the relationship of the <span class="hlt">CAI</span> ratings with objectively measured cognition, functional capacity, and role functioning. Comparison of schizophrenia patients with healthy controls further validates the <span class="hlt">CAI</span> as an co-primary measure of cognitive deficits. Also, <span class="hlt">CAI</span> change scores were strongly related to objective cognitive change indicating sensitivity to change. Copyright © 2016 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012M%26PS...47.1062R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012M%26PS...47.1062R"><span>A new model for the origin of Type-B and Fluffy Type-A <span class="hlt">CAIs</span>: Analogies to remelted compound chondrules</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Rubin, Alan E.</p> <p>2012-06-01</p> <p>In the scenario developed here, most types of calcium-aluminum-rich inclusions (<span class="hlt">CAIs</span>) formed near the Sun where they developed Wark-Lovering rims before being transported by aerodynamic forces throughout the nebula. The amount of ambient dust in the nebula varied with heliocentric distance, peaking in the CV-CK formation location. Literature data show that accretionary rims (which occur outside the Wark-Lovering rims) around <span class="hlt">CAIs</span> contain substantial 16O-rich forsterite, suggesting that, at this time, the ambient dust in the nebula consisted largely of 16O-rich forsterite. Individual sub-millimeter-size Compact Type-A <span class="hlt">CAIs</span> (each surrounded by a Wark-Lovering rim) collided in the CV-CK region and stuck together (in a manner similar to that of sibling compound chondrules); the CTAs were mixed with small amounts of 16O-rich mafic dust and formed centimeter-size compound objects (large Fluffy Type-A <span class="hlt">CAIs</span>) after experiencing minor melting. In contrast to other types of <span class="hlt">CAIs</span>, centimeter-size Type-B <span class="hlt">CAIs</span> formed directly in the CV-CK region after gehlenite-rich Compact Type-A <span class="hlt">CAIs</span> collided and stuck together, incorporated significant amounts of 16O-rich forsteritic dust (on the order of 10-15%) and probably some anorthite, and experienced extensive melting and partial evaporation. (Enveloping compound chondrules formed in an analogous manner.) In those cases where appreciably higher amounts of 16O-rich forsterite (on the order of 25%) (and perhaps minor anorthite and pyroxene) were incorporated into compound Type-A objects prior to melting, centimeter-size forsterite-bearing Type-B <span class="hlt">CAIs</span> (B3 inclusions) were produced. Type-B1 inclusions formed from B<span class="hlt">2</span> inclusions that collided with and stuck to melilite-rich Compact Type-A <span class="hlt">CAIs</span> and experienced high-temperature processing.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://files.eric.ed.gov/fulltext/ED265843.pdf','ERIC'); return false;" href="http://files.eric.ed.gov/fulltext/ED265843.pdf"><span>Micro-<span class="hlt">CAI</span> in Education: Some Considerations.</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>Majsterek, David</p> <p></p> <p>This paper focuses on the applications which best suit the microcomputer in an educational setting with emphasis on adapting effective pedagogical practice to the computer's programability and delivery capabilities. Discovery learning and "being told" are identified as two types of computer assisted instruction (<span class="hlt">CAI</span>) and sample uses of…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2005ksce.book..331M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2005ksce.book..331M"><span><span class="hlt">CAI</span> System with Multi-Media Text Through Web Browser for NC Lathe Programming</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mizugaki, Yoshio; Kikkawa, Koichi; Mizui, Masahiko; Kamijo, Keisuke</p> <p></p> <p>A new Computer Aided Instruction (<span class="hlt">CAI</span>) system for NC lathe programming has been developed with use of multi-media texts including movies, animations, pictures, sound and texts through Web browser. Although many <span class="hlt">CAI</span> systems developed previously for NC programming consist of text-based instructions, it is difficult for beginners to learn NC programming with use of them. In the developed <span class="hlt">CAI</span> system, multi-media texts are adopted for the help of users' understanding, and it is available through Web browser anytime and anywhere. Also the error log is automatically recorded for the future references. According to the NC programming coded by a user, the movement of the NC lathe is animated and shown in the monitor screen in front of the user. If its movement causes the collision between a cutting tool and the lathe, some sound and the caution remark are generated. If the user makes mistakes some times at a certain stage in learning NC, the corresponding suggestion is shown in the form of movies, animations, and so forth. By using the multimedia texts, users' attention is kept concentrated during a training course. In this paper, the configuration of the <span class="hlt">CAI</span> system is explained and the actual procedures for users to learn the NC programming are also explained too. Some beginners tested this <span class="hlt">CAI</span> system and their results are illustrated and discussed from the viewpoint of the efficiency and usefulness of this <span class="hlt">CAI</span> system. A brief conclusion is also mentioned.</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://hdl.handle.net/2060/19940011882','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19940011882"><span>INAA of <span class="hlt">CAIs</span> from the Maralinga CK4 chondrite: Effects of parent body thermal metamorphism</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Lindstrom, D. J.; Keller, L. P.; Martinez, R. R.</p> <p>1993-01-01</p> <p>Maralinga is an anomalous CK4 carbonaceous chondrite which contains numerous Ca-, Al-rich inclusions (<span class="hlt">CAI</span>'s) unlike the other members of the CK group. These <span class="hlt">CAI</span>'s are characterized by abundant green hercynitic spinel intergrown with plagioclase and high-Ca clinopyroxene, and a total lack of melilite. Instrumental Neutron Activation Analysis (INAA) was used to further characterize the meteorite, with special focus on the <span class="hlt">CAI</span>'s. High sensitivity INAA was done on eight sample disks about 100-150 microns in diameter obtained from a normal 30 micron thin section with a diamond microcoring device. The <span class="hlt">CAI</span>'s are enriched by 60-70X bulk meteorite values in Zn, suggesting that the substantial exchange of Fe for Mg that made the spinel in the <span class="hlt">CAI</span>'s hercynitic also allowed efficient scavenging of Zn from the rest of the meteorite during parent body thermal metamorphism. Less mobile elements appear to have maintained their initial heterogeneity.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20030110823&hterms=diversity&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3Ddiversity','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20030110823&hterms=diversity&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3Ddiversity"><span>Al-rich Chondrules: Petrologic Basis for Their Diversity, and Relation to Type C <span class="hlt">CAIs</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>MacPherson, G. J.; Huss, G. R.</p> <p>2003-01-01</p> <p>Al-rich chondrules share mineralogical and chemical properties with, and are intermediate in a volatility sense between, <span class="hlt">CAIs</span> and ferromagnesian chondrules. In some way they must be petrogenetic links between the two. A recent upsurge of interest in Al-rich chondrules is due to their constituent plagioclase feldspar and Al-rich glass being amenable to successful ion microprobe searches for radiogenic Mg-26, the decay product of Al-26 (t(sub 1/<span class="hlt">2</span>) = 720,000 y). This has allowed estimates to be made of the time duration between <span class="hlt">CAI</span> formation and the onset of Al-rich (and possibly, by extension, ferromagnesian) chondrule formation, on the order of 1.5-<span class="hlt">2</span>.5 million years.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/12163501','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/12163501"><span><span class="hlt">Cai</span>T of Escherichia coli, a new transporter catalyzing L-carnitine/gamma -butyrobetaine exchange.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Jung, Heinrich; Buchholz, Marion; Clausen, Jurgen; Nietschke, Monika; Revermann, Anne; Schmid, Roland; Jung, Kirsten</p> <p>2002-10-18</p> <p>l-Carnitine is essential for beta-oxidation of fatty acids in mitochondria. Bacterial metabolic pathways are used for the production of this medically important compound. Here, we report the first detailed functional characterization of the <span class="hlt">cai</span>T gene product, a putative transport protein whose function is required for l-carnitine conversion in Escherichia coli. The <span class="hlt">cai</span>T gene was overexpressed in E. coli, and the gene product was purified by affinity chromatography and reconstituted into proteoliposomes. Functional analyses with intact cells and proteoliposomes demonstrated that <span class="hlt">Cai</span>T is able to catalyze the exchange of l-carnitine for gamma-butyrobetaine, the excreted end product of l-carnitine conversion in E. coli, and related betaines. Electrochemical ion gradients did not significantly stimulate l-carnitine uptake. Analysis of l-carnitine counterflow yielded an apparent external K(m) of 105 microm and a turnover number of 5.5 s(-1). Contrary to related proteins, <span class="hlt">Cai</span>T activity was not modulated by osmotic stress. l-Carnitine binding to <span class="hlt">Cai</span>T increased the protein fluorescence and caused a red shift in the emission maximum, an observation explained by ligand-induced conformational alterations. The fluorescence effect was specific for betaine structures, for which the distance between trimethylammonium and carboxyl groups proved to be crucial for affinity. Taken together, the results suggest that <span class="hlt">Cai</span>T functions as an exchanger (antiporter) for l-carnitine and gamma-butyrobetaine according to the substrate/product antiport principle.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JCrGr.486..162S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JCrGr.486..162S"><span>Crystal Growth and Scintillation Properties of Eu<span class="hlt">2</span>+ doped Cs4<span class="hlt">CaI</span>6 and Cs4SrI6</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Stand, L.; Zhuravleva, M.; Chakoumakos, B.; Johnson, J.; Loyd, M.; Wu, Y.; Koschan, M.; Melcher, C. L.</p> <p>2018-03-01</p> <p>In this work we present the crystal growth and scintillation properties of two new ternarymetal halide scintillators activated with divalent europium, Cs4<span class="hlt">CaI</span>6 and Cs4SrI6. Single crystals of each compound were grown in evacuated quartz ampoules via the vertical Bridgman technique using a two-zone transparent furnace. Single crystal X-ray diffraction experiments showed that both crystals have a trigonal (R-3c) structure, with a density of 3.99 g/cm3 and 4.03 g/cm3. The radioluminescence and photoluminescence measurements showed typical luminescence properties due to the 5d-4f radiative transitions in Eu<span class="hlt">2</span>+. At this early stage of development Cs4SrI6:Eu and Cs4<span class="hlt">CaI</span>6:Eu have shown very promising scintillation properties, with light yields and energy resolutions of 62,300 ph/MeV and 3.3%, and 51,800 photons/MeV and 3.6% at 662 keV, respectively.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20140010652','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20140010652"><span>A FIB/TEM/Nanosims Study of a Wark-Lovering Rim on an Allende <span class="hlt">CAI</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Keller, L. P.; Needham, A. W.; Messenger, S.</p> <p>2014-01-01</p> <p>Ca- Al-rich inclusions (<span class="hlt">CAIs</span>) are commonly surrounded by Wark-Lovering (WL) rims - thin (approx. 50 micrometers) multilayered sequences - whose mineralogy is dominated by high temperature minerals similar to those that occur in the cores of <span class="hlt">CAIs</span> [1]. The origins of these WL rims involved high temperature events in the early nebula such as condensation, flashheating or reaction with a nebular reservoir, or combinations of these processes. These rims formed after <span class="hlt">CAI</span> formation but prior to accretion into their parent bodies. We have undertaken a coordinated mineralogical and isotopic study of WL rims to determine the formation conditions of the individual layers and to constrain the isotopic reservoirs they interacted with during their history. We focus here on the spinel layer, the first-formed highest- temperature layer in the WL rim sequence. Results and Discussion: We have performed mineralogical, chemical and isotopic analyses of an unusual ultrarefractory inclusion from the Allende CV3 chondrite (SHAL) consisting of an approx. 500 micrometers long single crystal of hibonite and co-existing coarsegrained perovskite. SHAL is partially surrounded by WL rim. We previously reported on the mineralogy, isotopic compositions and trace elements in SHAL [<span class="hlt">2</span>-4]. The spinel layer in the WL rim is present only on the hibonite and terminates abruptly at the contact with the coarse perovskite. This simple observation shows that the spinel layer is not a condensate in this case (otherwise spinel would have condensed on the perovskite as well). The spinel layer appears to have formed by gas-phase corrosion of the hibonite by Mg-rich vapors such that the spinel layer grew at the expense of the hibonite. We also found that the spinel layer has the same 16Orich composition as the hibonite. The spinel layer is polycrystalline and individual crystals do not show a crystallographic relationship with the hibonite. An Al-diopside layer overlies the spinel layer, and is present on both</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://eric.ed.gov/?q=red+AND+wings&id=ED189125','ERIC'); return false;" href="https://eric.ed.gov/?q=red+AND+wings&id=ED189125"><span>Evaluation of Title I <span class="hlt">CAI</span> Programs at Minnesota State Correctional Institutions.</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>Sandman, Richard S.; Welch, Wayne W.</p> <p></p> <p>Three Minnesota correctional institutions used computer-assisted instruction (<span class="hlt">CAI</span>) on PLATO terminals to improve reading and mathematics skills: (1) the State Reformatory for Men, St. Cloud (males, ages 17-21); (<span class="hlt">2</span>) the Minnesota Home School, Sauk Centre (males and females, ages 12-18); and (3) the State Training School, Red Wing (males, ages…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2003EAEJA.....9292C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2003EAEJA.....9292C"><span>B and Mg isotopic variations in Leoville mrs-06 type B1 <span class="hlt">cai</span>:origin of 10Be and 26Al</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Chaussidon, M.; Robert, F.; Russel, S. S.; Gounelle, M.; Ash, R. D.</p> <p>2003-04-01</p> <p>The finding [1-3] in Ca-Al-rich refractory inclusions (<span class="hlt">CAI</span>) of primitive chondrites of traces of the in situ decay of radioactive 10Be (half-life 1.5Myr) indicates that irradiation of the protosolar nebula by the young Sun in its T-Tauri phase has produced significant amounts of the Li-Be-B elements. This irradiation may have produced also some or all of the short-lived 26Al (half-life 0.7Myr) and 41Ca (half-life 0.1Myr) previously detected in <span class="hlt">CAIs</span>. To constrain the origin of 10Be and 10Al it is important to look for coupled variations in the 10Be/9Be and 26Al/27Al ratios in <span class="hlt">CAIs</span> and to understand the processes responsible for these variations (e.g. variations in the fluences of irradiation, secondary perturbations of the <span class="hlt">CAIs</span>, ...) We have thus studied the Li and B isotopic compositions and the Be/Li and Be/B concentration ratios in one <span class="hlt">CAI</span> (MRS-06) from the Leoville CV3 chondrite in which large variations of the Mg isotopic compositions showing both the in situ decay of 26Al and the secondary redistribution of Mg isotopes have been observed [4]. The results show large variations for the Li and B isotopic compositions (^7Li/^6Li ranging from 11.02±0.21 to 11.82±0.07, and 10B/11B ratios ranging from 0.2457±0.0053 to 0.2980±0.0085). The ^7Li/^6Li ratio tend to decrease towards the rim of the inclusion. The 10B/11B ratios are positively correlated with the ^9Be/11B ratios indicating the in situ decay of 10Be. However perturbations of the 10Be/B system are observed. They would correspond to an event which occurred approximately <span class="hlt">2</span>Myr after the formation of the <span class="hlt">CAI</span> and the irradiation of the <span class="hlt">CAI</span> precursors which is responsible for the 10Be observed in the core of the <span class="hlt">CAI</span>. These perturbations seem compatible with those observed for the 26Al/Mg system but they might be due to an irradiation of the already-formed, isolated <span class="hlt">CAI</span> which would have resulted in increased 10Be/^9Be ratios and low ^7Li/^6Li ratios in the margin of the <span class="hlt">CAI</span>. [1] McKeegan K. D. et al. (2000</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2009AGUFM.P11C1240D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009AGUFM.P11C1240D"><span>Characterizing Pyroxene Reaction Space in Calcium-Aluminum Rich Inclusions: Oxidation During <span class="hlt">CAI</span> Rim Formation</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Dyl, K. A.; Young, E. D.</p> <p>2009-12-01</p> <p>We define the reaction space that controls changes in pyroxene composition in <span class="hlt">CAIs</span> and Wark-Lovering (WL) rims in an oxidizing solar nebula. Ti-rich pyroxenes in <span class="hlt">CAIs</span> record a sub-solar oxygen fugacity (Ti3+/Ti4+~1.5). WL rim pyroxenes in the <span class="hlt">CAI</span> Leoville 144A have a distinctly lower oxidation state.This difference supports WL rim condensation in an environment of increasing O<span class="hlt">2</span>(g) and Mg(g) (Simon et al. 2005). We used the following phase components to identify four linearly independent reactions (Thompson 1982): diopside, CaTs (Al<span class="hlt">2</span>Mg-1Si-1), T3 (Ti3+AlMg-1Si-1), T4 (Ti4+Al<span class="hlt">2</span>Mg-1Si-<span class="hlt">2</span>), En (MgCa-1), perovskite, O(g), Mg(g), SiO(g), and Ca(g). Compositional variation in this system is dominated by two reactions. The first is oxidation of Ti3+ via reaction with O and Mg in the gas phase: 1.5 O(g) + Mg(g) → ¼ Di + [Ti4+Mg3/4Ti3+-1Ca-1/4Si-1/<span class="hlt">2</span>] (1). Pyroxene is produced and En is introduced. The second reaction (<span class="hlt">2</span>) is perovskite formation. It is observed in the WL rim of Leoville 144A, and experiments confirm that an elevated Ti component converts pyroxene to perovskite(Gupta et al. 1973). MgCa-1 is the third linearly independent reaction (3). They combine to give: ½ Di + x Ca(g)→ x Mg(g)+ Pv + [Mg1/<span class="hlt">2</span>-xSiTi4+-1Ca-1/<span class="hlt">2</span>+x](<span class="hlt">2</span>,3). Unlike (1), pyroxene is consumed in this reaction. The parameter x defines the extent of Mg-Ca exchange. When x > 0.5, WL rim formation occurs in an environment where Mg is volatile and Ca condenses. The reaction space defined by reactions (1) and (<span class="hlt">2</span>,3) describes the transition from <span class="hlt">CAI</span> interior to WL rims. WL rim pyroxene Ti contents, [CaTs], and Ca < 1 pfu are all explained in this space. The fourth linearly independent reaction is SiO(g):1/8 Di + ¼ Mg(g)→ ¾ SiO(g) + [Mg3/8Ca1/8Ti4+Ti3+-1Si-1/<span class="hlt">2</span>](4). Silica reduction forms Ti4+, releasing SiO(g). (4) does not describe the oxidation of Ti3+ in WL rim pyroxene, but (1) - (4) results in En formation directly from the gas phase. This may explain WL rim analyses that have Si contents in excess</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://files.eric.ed.gov/fulltext/ED152294.pdf','ERIC'); return false;" href="http://files.eric.ed.gov/fulltext/ED152294.pdf"><span>The Relevance of AI Research to <span class="hlt">CAI</span>.</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>Kearsley, Greg P.</p> <p></p> <p>This article provides a tutorial introduction to Artificial Intelligence (AI) research for those involved in Computer Assisted Instruction (<span class="hlt">CAI</span>). The general theme is that much of the current work in AI, particularly in the areas of natural language understanding systems, rule induction, programming languages, and socratic systems, has important…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://files.eric.ed.gov/fulltext/ED058725.pdf','ERIC'); return false;" href="http://files.eric.ed.gov/fulltext/ED058725.pdf"><span>Who Should Develop Instructional Materials for <span class="hlt">CAI</span>?</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>Seidel, Robert J.</p> <p></p> <p>The nonprofit special organization as a developer of computer-administered instruction (<span class="hlt">CAI</span>) is advocated in this paper. The organization of universities and their mode of operation do not lend themselves to instructional product development. Faculty members engage in such efforts on a part-time basis and in competition with higher priority…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26133743','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26133743"><span>Gender Role, Gender Identity and Sexual Orientation in <span class="hlt">CAIS</span> ("XY-Women") Compared With Subfertile and Infertile 46,XX Women.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Brunner, Franziska; Fliegner, Maike; Krupp, Kerstin; Rall, Katharina; Brucker, Sara; Richter-Appelt, Hertha</p> <p>2016-01-01</p> <p>The perception of gender development of individuals with complete androgen insensitivity syndrome (<span class="hlt">CAIS</span>) as unambiguously female has recently been challenged in both qualitative data and case reports of male gender identity. The aim of the mixed-method study presented was to examine the self-perception of <span class="hlt">CAIS</span> individuals regarding different aspects of gender and to identify commonalities and differences in comparison with subfertile and infertile XX-chromosomal women with diagnoses of Mayer-Rokitansky-Küster-Hauser syndrome (MRKHS) and polycystic ovary syndrome (PCOS). The study sample comprised 11 participants with <span class="hlt">CAIS</span>, 49 with MRKHS, and 55 with PCOS. Gender identity was assessed by means of a multidimensional instrument, which showed significant differences between the <span class="hlt">CAIS</span> group and the XX-chromosomal women. Other-than-female gender roles and neither-female-nor-male sexes/genders were reported only by individuals with <span class="hlt">CAIS</span>. The percentage with a not exclusively androphile sexual orientation was unexceptionally high in the <span class="hlt">CAIS</span> group compared to the prevalence in "normative" women and the clinical groups. The findings support the assumption made by Meyer-Bahlburg ( 2010 ) that gender outcome in people with <span class="hlt">CAIS</span> is more variable than generally stated. Parents and professionals should thus be open to courses of gender development other than typically female in individuals with <span class="hlt">CAIS</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.dtic.mil/docs/citations/ADA186080','DTIC-ST'); return false;" href="http://www.dtic.mil/docs/citations/ADA186080"><span>A Prototype of Pilot Knowledge Evaluation by an Intelligent <span class="hlt">CAI</span> (Computer -Aided Instruction) System Using a Bayesian Diagnostic Model.</span></a></p> <p><a target="_blank" href="http://www.dtic.mil/">DTIC Science & Technology</a></p> <p></p> <p>1987-06-01</p> <p>to a field of research called Computer-Aided Instruction (<span class="hlt">CAI</span>). <span class="hlt">CAI</span> is a powerful methodology for enhancing the overall quaiity and effectiveness of...provides a very powerful tool for statistical inference, especially when pooling informations from different source is appropriate. Thus. prior...04 , <span class="hlt">2</span> ’ .. ."k, + ++ ,,;-+-,..,,..v ->’,0,,.’ I The power of the model lies in its ability to adapt a diagnostic session to the level of knowledge</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19850054072&hterms=Prize&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DTitle%26N%3D0%26No%3D10%26Ntt%3DPrize','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19850054072&hterms=Prize&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DTitle%26N%3D0%26No%3D10%26Ntt%3DPrize"><span>Willy: A prize noble Ur-Fremdling - Its history and implications for the formation of Fremdlinge and <span class="hlt">CAI</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Armstrong, J. T.; El Goresy, A.; Wasserburg, G. J.</p> <p>1985-01-01</p> <p>The structure and composition of Willy, a 150-micron-diameter Fremdling in <span class="hlt">CAI</span> 5241 from the Allende meteorite, are investigated using optical, secondary-electron, and electron-backscatter microscopy and electron-microprobe analysis. The results are presented in diagrams, maps, tables, graphs, and micrographs and compared with those for other Allende Fremdlinge. Willy is found to have a concentric-zone structure comprising a complex porous core of magnetite, metal, sulfide, scheelite, and other minor phases; a compact magnetite-apatite mantle; a thin (20 microns or less) reaction-assemblage zone; and a dense outer rim of fassaite with minor spinel. A multistage formation sequence involving changes in T and fO<span class="hlt">2</span> and preceding the introduction of Willy into the <span class="hlt">CAI</span> (which itself preceded <span class="hlt">CAI</span> spinel and silicate formation) is postulated, and it is inferred from the apparent lack of post-capture recrystallization that Willy has not been subjected to temperatures in excess of 600 C and may represent the precursor material for many other Fremdlinge.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20140001393','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20140001393"><span>A FIB/TEM Study of a Complex Wark-Lovering Rim on a Vigarano <span class="hlt">CAI</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Keller, L. P.; Needham, A. W.; Messenger, S.</p> <p>2013-01-01</p> <p>Wark-Lovering (WL) rims are thin multilayered mineral sequences that surround most Ca, Al-rich inclusions (<span class="hlt">CAIs</span>). Several processes have been proposed for WL rim formation, including condensation, flash-heating or reaction with a nebular reservoir, or combinations of these [e.g. 1-7], but no consensus exists. Our previous coordinated transmission electron microscope (TEM) and NanoSIMS O isotopic measurements showed that a WL rim experienced flash heating events in a nebular environment with planetary O isotopic composition, distinct from the (16)O-rich formation environment [6]. Our efforts have focused on <span class="hlt">CAIs</span> from the CV(sub red) chondrites, especially Vigarano, because these have escaped much of the parent body alteration effects that are common in <span class="hlt">CAIs</span> from CV(sub ox) group.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://files.eric.ed.gov/fulltext/ED186017.pdf','ERIC'); return false;" href="http://files.eric.ed.gov/fulltext/ED186017.pdf"><span>Low-Cost Computer-Aided Instruction/Computer-Managed Instruction (<span class="hlt">CAI</span>/CMI) System: Feasibility Study. Final Report.</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>Lintz, Larry M.; And Others</p> <p></p> <p>This study investigated the feasibility of a low cost computer-aided instruction/computer-managed instruction (<span class="hlt">CAI</span>/CMI) system. Air Force instructors and training supervisors were surveyed to determine the potential payoffs of various <span class="hlt">CAI</span> and CMI functions. Results indicated that a wide range of capabilities had potential for resident technical…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://files.eric.ed.gov/fulltext/ED153600.pdf','ERIC'); return false;" href="http://files.eric.ed.gov/fulltext/ED153600.pdf"><span>The Cost of <span class="hlt">CAI</span>: A Matter of Assumptions.</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>Kearsley, Greg P.</p> <p></p> <p>Cost estimates for Computer Assisted Instruction (<span class="hlt">CAI</span>) depend crucially upon the particular assumptions made about the components of the system to be included in the costs, the expected lifetime of the system and courseware, and the anticipated student utilization of the system/courseware. The cost estimates of three currently operational systems…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.P51A2558C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.P51A2558C"><span>Exploring Chondrule and <span class="hlt">CAI</span> Rims Using Micro- and Nano-Scale Petrological and Compositional Analysis</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Cartwright, J. A.; Perez-Huerta, A.; Leitner, J.; Vollmer, C.</p> <p>2017-12-01</p> <p>As the major components within chondrites, chondrules (mm-sized droplets of quenched silicate melt) and calcium-aluminum-rich inclusions (<span class="hlt">CAI</span>, refractory) represent the most abundant and the earliest materials that solidified from the solar nebula. However, the exact formation mechanisms of these clasts, and whether these processes are related, remains unconstrained, despite extensive petrological and compositional study. By taking advantage of recent advances in nano-scale tomographical techniques, we have undertaken a combined micro- and nano-scale study of <span class="hlt">CAI</span> and chondrule rim morphologies, to investigate their formation mechanisms. The target lithologies for this research are Wark-Lovering rims (WLR), and fine-grained rims (FGR) around <span class="hlt">CAIs</span> and chondrules respectively, present within many chondrites. The FGRs, which are up to 100 µm thick, are of particular interest as recent studies have identified presolar grains within them. These grains predate the formation of our Solar System, suggesting FGR formation under nebular conditions. By contrast, WLRs are 10-20 µm thick, made of different compositional layers, and likely formed by flash-heating shortly after <span class="hlt">CAI</span> formation, thus recording nebular conditions. A detailed multi-scale study of these respective rims will enable us to better understand their formation histories and determine the potential for commonality between these two phases, despite reports of an observed formation age difference of up to <span class="hlt">2</span>-3 Myr. We are using a combination of complimentary techniques on our selected target areas: 1) Micro-scale characterization using standard microscopic and compositional techniques (SEM-EBSD, EMPA); <span class="hlt">2</span>) Nano-scale characterization of structures using transmission electron microscopy (TEM) and elemental, isotopic and tomographic analysis with NanoSIMS and atom probe tomography (APT). Preliminary nano-scale APT analysis of FGR morphologies within the Allende carbonaceous chondrite has successfully discerned</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20140010679','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20140010679"><span>Microstructures of Hibonite From an ALH A77307 (CO3.0) <span class="hlt">CAI</span>: Evidence for Evaporative Loss of Calcium</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Han, Jangmi; Brearley, Adrian J.; Keller, Lindsay P.</p> <p>2014-01-01</p> <p>Hibonite is a comparatively rare, primary phase found in some <span class="hlt">CAIs</span> from different chondrite groups and is also common in Wark-Lovering rims [1]. Hibonite is predicted to be one of the earliest refractory phases to form by equilibrium condensation from a cooling gas of solar composition [<span class="hlt">2</span>] and, therefore, can be a potential recorder of very early solar system processes. In this study, we describe the microstructures of hibonite from one <span class="hlt">CAI</span> in ALH A77307 (CO3.0) using FIB/TEM techniques in order to reconstruct its formational history.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/20542412','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/20542412"><span>The Cognitive Assessment Interview (<span class="hlt">CAI</span>): development and validation of an empirically derived, brief interview-based measure of cognition.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Ventura, Joseph; Reise, Steven P; Keefe, Richard S E; Baade, Lyle E; Gold, James M; Green, Michael F; Kern, Robert S; Mesholam-Gately, Raquelle; Nuechterlein, Keith H; Seidman, Larry J; Bilder, Robert M</p> <p>2010-08-01</p> <p>Practical, reliable "real world" measures of cognition are needed to supplement neurocognitive performance data to evaluate possible efficacy of new drugs targeting cognitive deficits associated with schizophrenia. Because interview-based measures of cognition offer one possible approach, data from the MATRICS initiative (n=176) were used to examine the psychometric properties of the Schizophrenia Cognition Rating Scale (SCoRS) and the Clinical Global Impression of Cognition in Schizophrenia (CGI-CogS). We used classical test theory methods and item response theory to derive the 10-item Cognitive Assessment Interview (<span class="hlt">CAI</span>) from the SCoRS and CGI-CogS ("parent instruments"). Sources of information for <span class="hlt">CAI</span> ratings included the patient and an informant. Validity analyses examined the relationship between the <span class="hlt">CAI</span> and objective measures of cognitive functioning, intermediate measures of cognition, and functional outcome. The rater's score from the newly derived <span class="hlt">CAI</span> (10 items) correlate highly (r=.87) with those from the combined set of the SCoRS and CGI-CogS (41 items). Both the patient (r=.82) and the informant (r=.95) data were highly correlated with the rater's score. The <span class="hlt">CAI</span> was modestly correlated with objectively measured neurocognition (r=-.32), functional capacity (r=-.44), and functional outcome (r=-.32), which was comparable to the parent instruments. The <span class="hlt">CAI</span> allows for expert judgment in evaluating a patient's cognitive functioning and was modestly correlated with neurocognitive functioning, functional capacity, and functional outcome. The <span class="hlt">CAI</span> is a brief, repeatable, and potentially valuable tool for rating cognition in schizophrenia patients who are participating in clinical trials. Copyright 2010 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017GeCoA.201....6P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017GeCoA.201....6P"><span>Calcium-aluminum-rich inclusions with fractionation and unidentified nuclear effects (FUN <span class="hlt">CAIs</span>): II. Heterogeneities of magnesium isotopes and 26Al in the early Solar System inferred from in situ high-precision magnesium-isotope measurements</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Park, Changkun; Nagashima, Kazuhide; Krot, Alexander N.; Huss, Gary R.; Davis, Andrew M.; Bizzarro, Martin</p> <p>2017-03-01</p> <p>Calcium-aluminum-rich inclusions with isotopic mass fractionation effects and unidentified nuclear isotopic anomalies (FUN <span class="hlt">CAIs</span>) have been studied for more than 40 years, but their origins remain enigmatic. Here we report in situ high precision measurements of aluminum-magnesium isotope systematics of FUN <span class="hlt">CAIs</span> by secondary ion mass spectrometry (SIMS). Individual minerals were analyzed in six FUN <span class="hlt">CAIs</span> from the oxidized CV3 carbonaceous chondrites Axtell (compact Type A <span class="hlt">CAI</span> Axtell 2271) and Allende (Type B <span class="hlt">CAIs</span> C1 and EK1-4-1, and forsterite-bearing Type B <span class="hlt">CAIs</span> BG82DH8, CG-14, and TE). Most of these <span class="hlt">CAIs</span> show evidence for excess 26Mg due to the decay of 26Al. The inferred initial 26Al/27Al ratios [(26Al/27Al)0] and the initial magnesium isotopic compositions (δ26Mg0) calculated using an exponential law with an exponent β of 0.5128 are (3.1 ± 1.6) × 10-6 and 0.60 ± 0.10‰ (Axtell 2271), (3.7 ± 1.5) × 10-6 and -0.20 ± 0.05‰ (BG82DH8), (<span class="hlt">2.2</span> ± 1.1) × 10-6 and -0.18 ± 0.05‰ (C1), (<span class="hlt">2</span>.3 ± <span class="hlt">2</span>.4) × 10-5 and -<span class="hlt">2</span>.23 ± 0.37‰ (EK1-4-1), (1.5 ± 1.1) × 10-5 and -0.42 ± 0.08‰ (CG-14), and (5.3 ± 0.9) × 10-5 and -0.05 ± 0.08‰ (TE) with <span class="hlt">2</span>σ uncertainties. We infer that FUN <span class="hlt">CAIs</span> recorded heterogeneities of magnesium isotopes and 26Al in the <span class="hlt">CAI</span>-forming region(s). Comparison of 26Al-26Mg systematics, stable isotope (oxygen, magnesium, calcium, and titanium) and trace element studies of FUN and non-FUN igneous <span class="hlt">CAIs</span> indicates that there is a continuum among these <span class="hlt">CAI</span> types. Based on these observations and evaporation experiments on <span class="hlt">CAI</span>-like melts, we propose a generic scenario for the origin of igneous (FUN and non-FUN) <span class="hlt">CAIs</span>: (i) condensation of isotopically normal solids in an 16O-rich gas of approximately solar composition; (ii) formation of <span class="hlt">CAI</span> precursors by aggregation of these solids together with variable abundances of isotopically anomalous grains-possible carriers of unidentified nuclear (UN) effects; and (iii) melt evaporation of these precursors</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('http://adsabs.harvard.edu/abs/2015GeCoA.153..183F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015GeCoA.153..183F"><span>Evidence for an early nitrogen isotopic evolution in the solar nebula from volatile analyses of a <span class="hlt">CAI</span> from the CV3 chondrite NWA 8616</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Füri, Evelyn; Chaussidon, Marc; Marty, Bernard</p> <p>2015-03-01</p> <p>Nitrogen and noble gas (Ne-Ar) abundances and isotope ratios, determined by CO<span class="hlt">2</span> laser extraction static mass spectrometry analysis, as well as Al-Mg and O isotope data from secondary ion mass spectrometry (SIMS) analyses, are reported for a type B calcium-aluminum-rich inclusion (<span class="hlt">CAI</span>) from the CV3 chondrite NWA 8616. The high (26Al/27Al)i ratio of (5.06 ± 0.50) × 10-5 dates the last melting event of the <span class="hlt">CAI</span> at 39-99+109ka after "time zero", limiting the period during which high-temperature exchanges between the <span class="hlt">CAI</span> and the nebular gas could have occurred to a very short time interval. Partial isotopic exchange with a 16O-poor reservoir resulted in Δ17O > -5‰ for melilite and anorthite, whereas spinel and Al-Ti-pyroxene retain the inferred original 16O-rich signature of the solar nebula (Δ17O ⩽ -20‰). The low 20Ne/22Ne (⩽0.83) and 36Ar/38Ar (⩽0.75) ratios of the <span class="hlt">CAI</span> rule out the presence of any trapped planetary or solar noble gases. Cosmogenic 21Ne and 38Ar abundances are consistent with a cosmic ray exposure (CRE) age of ∼14 to 20 Ma, assuming CR fluxes similar to modern ones, without any evidence for pre-irradiation of the <span class="hlt">CAI</span> before incorporation into the meteorite parent body. Strikingly, the <span class="hlt">CAI</span> contains 1.4-3.4 ppm N with a δ15N value of +8‰ to +30‰. Even after correcting the measured δ15N values for cosmogenic 15N produced in situ, the <span class="hlt">CAI</span> is highly enriched in 15N compared to the protosolar nebula (δ15NPSN = -383 ± 8‰; Marty et al., 2011), implying that the <span class="hlt">CAI</span>-forming region was contaminated by 15N-rich material within the first 0.15 Ma of Solar System history, or, alternatively, that the <span class="hlt">CAI</span> was ejected into the outer Solar System where it interacted with a 15N-rich reservoir.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017GeCoA.201...25W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017GeCoA.201...25W"><span>Thermal and chemical evolution in the early solar system as recorded by FUN <span class="hlt">CAIs</span>: Part I - Petrology, mineral chemistry, and isotopic composition of Allende FUN <span class="hlt">CAI</span> CMS-1</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Williams, C. D.; Ushikubo, T.; Bullock, E. S.; Janney, P. E.; Hines, R. R.; Kita, N. T.; Hervig, R. L.; MacPherson, G. J.; Mendybaev, R. A.; Richter, F. M.; Wadhwa, M.</p> <p>2017-03-01</p> <p>Detailed petrologic, geochemical and isotopic analyses of a new FUN <span class="hlt">CAI</span> from the Allende CV3 meteorite (designated CMS-1) indicate that it formed by extensive melting and evaporation of primitive precursor material(s). The precursor material(s) condensed in a 16O-rich region (δ17O and δ18O ∼ -49‰) of the inner solar nebula dominated by gas of solar composition at total pressures of ∼10-3-10-6 bar. Subsequent melting of the precursor material(s) was accompanied by evaporative loss of magnesium, silicon and oxygen resulting in large mass-dependent isotope fractionations in these elements (δ25Mg = 30.71-39.26‰, δ29Si = 14.98-16.65‰, and δ18O = -41.57 to -15.50‰). This evaporative loss resulted in a bulk composition similar to that of compact Type A and Type B <span class="hlt">CAIs</span>, but very distinct from the composition of the original precursor condensate(s). Kinetic fractionation factors and the measured mass-dependent fractionation of silicon and magnesium in CMS-1 suggest that ∼80% of the silicon and ∼85% of the magnesium were lost from its precursor material(s) through evaporative processes. These results suggest that the precursor material(s) of normal and FUN <span class="hlt">CAIs</span> condensed in similar environments, but subsequently evolved under vastly different conditions such as total gas pressure. The chemical and isotopic differences between normal and FUN <span class="hlt">CAIs</span> could be explained by sorting of early solar system materials into distinct physical and chemical regimes, in conjunction with discrete heating events, within the protoplanetary disk.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3184638','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3184638"><span>The Cognitive Assessment Interview (<span class="hlt">CAI</span>): Development and Validation of an Empirically Derived, Brief Interview-Based Measure of Cognition</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Ventura, Joseph; Reise, Steven P.; Keefe, Richard S. E.; Baade, Lyle E.; Gold, James M.; Green, Michael F.; Kern, Robert S.; Mesholam-Gately, Raquelle; Nuechterlein, Keith H.; Seidman, Larry J.; Bilder, Robert M.</p> <p>2011-01-01</p> <p>Background Practical, reliable “real world” measures of cognition are needed to supplement neurocognitive performance data to evaluate possible efficacy of new drugs targeting cognitive deficits associated with schizophrenia. Because interview-based measures of cognition offer one possible approach, data from the MATRICS initiative (n=176) were used to examine the psychometric properties of the Schizophrenia Cognition Rating Scale (SCoRS) and the Clinical Global Impression of Cognition in Schizophrenia (CGI-CogS). Method We used classical test theory methods and item response theory to derive the 10 item Cognitive Assessment Interview (<span class="hlt">CAI</span>) from the SCoRS and CGI-Cogs (“parent instruments”). Sources of information for <span class="hlt">CAI</span> ratings included the patient and an informant. Validity analyses examined the relationship between the <span class="hlt">CAI</span> and objective measures of cognitive functioning, intermediate measures of cognition, and functional outcome. Results The rater’s score from the newly derived <span class="hlt">CAI</span> (10-items) correlate highly (r = .87) with those from the combined set of the SCoRS and CGI-CogS (41 items). Both the patient (r= .82) and the informant (r= .95) data were highly correlated with the rater’s score. The <span class="hlt">CAI</span> was modestly correlated with objectively measured neurocognition (r = −.32), functional capacity (r = −.44), and functional outcome (r = −.32), which was comparable to the parent instruments. Conclusions The <span class="hlt">CAI</span> allows for expert judgment in evaluating a patient’s cognitive functioning and was modestly correlated with neurocognitive functioning, functional capacity, and functional outcome. The <span class="hlt">CAI</span> is a brief, repeatable, and potentially valuable tool for rating cognition in schizophrenia patients who are participating in clinical trials. PMID:20542412</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017LPICo1987.6355F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017LPICo1987.6355F"><span>Oxygen, Magnesium, and Aluminum Isotopes in the Ivuna <span class="hlt">CAI</span>: Re-Examining High-Temperature Fractionations in CI Chondrites</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Frank, D. R.; Huss, G. R.; Nagashima, K.; Zolensky, M. E.; Le, L.</p> <p>2017-07-01</p> <p>The only whole <span class="hlt">CAI</span> preserved in the aqueously altered CI chondrites is 16O-rich and has no resolvable radiogenic Mg. Accretion of <span class="hlt">CAIs</span> by the CI parent object(s) may limit the precision of cosmochemical models that require a CI starting composition.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.dtic.mil/docs/citations/ADA239998','DTIC-ST'); return false;" href="http://www.dtic.mil/docs/citations/ADA239998"><span>A Design of Computer Aided Instructions (<span class="hlt">CAI</span>) for Undirected Graphs in the Discrete Math Tutorial (DMT). Part <span class="hlt">2</span></span></a></p> <p><a target="_blank" href="http://www.dtic.mil/">DTIC Science & Technology</a></p> <p></p> <p>1990-06-01</p> <p>The objective of this thesis research is to create a tutorial for teaching aspects of undirected graphs in discrete math . It is one of the submodules...of the Discrete Math Tutorial (DMT), which is a Computer Aided Instructional (<span class="hlt">CAI</span>) tool for teaching discrete math to the Naval Academy and the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016MS%26E..148a2083P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016MS%26E..148a2083P"><span>Numerical investigation of <span class="hlt">CAI</span> Combustion in the Opposed- Piston Engine with Direct and Indirect Water Injection</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Pyszczek, R.; Mazuro, P.; Teodorczyk, A.</p> <p>2016-09-01</p> <p>This paper is focused on the <span class="hlt">CAI</span> combustion control in a turbocharged <span class="hlt">2</span>-stroke Opposed-Piston (OP) engine. The barrel type OP engine arrangement is of particular interest for the authors because of its robust design, high mechanical efficiency and relatively easy incorporation of a Variable Compression Ratio (VCR). The other advantage of such design is that combustion chamber is formed between two moving pistons - there is no additional cylinder head to be cooled which directly results in an increased thermal efficiency. Furthermore, engine operation in a Controlled Auto-Ignition (<span class="hlt">CAI</span>) mode at high compression ratios (CR) raises a possibility of reaching even higher efficiencies and very low emissions. In order to control <span class="hlt">CAI</span> combustion such measures as VCR and water injection were considered for indirect ignition timing control. Numerical simulations of the scavenging and combustion processes were performed with the 3D CFD multipurpose AVL Fire solver. Numerous cases were calculated with different engine compression ratios and different amounts of directly and indirectly injected water. The influence of the VCR and water injection on the ignition timing and engine performance was determined and their application in the real engine was discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://files.eric.ed.gov/fulltext/ED078681.pdf','ERIC'); return false;" href="http://files.eric.ed.gov/fulltext/ED078681.pdf"><span>An Intelligent <span class="hlt">CAI</span> Monitor and Generative Tutor. Interim Report.</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>Koffman, Elliot B.; And Others</p> <p></p> <p>Design techniques for generative computer-assisted-instructional (<span class="hlt">CAI</span>) systems are described in this report. These are systems capable of generating problems for students and of deriving and monitoring solutions; problem difficulty, instructional pace, and depth of monitoring are all individually tailored and parts of the solution algorithms can…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://files.eric.ed.gov/fulltext/ED343582.pdf','ERIC'); return false;" href="http://files.eric.ed.gov/fulltext/ED343582.pdf"><span>The <span class="hlt">CAI</span>/Cooperative Learning Project. First Year Evaluation Report.</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>Beyer, Francine S.</p> <p></p> <p>This report presents a first year evaluation of the Computer Assisted Instruction (<span class="hlt">CAI</span>)/ Cooperative Learning Project, a 3-year collaborative effort by two Pennsylvania school districts--the Pittston Area School District and the Hatboro-Horsham School District--and Research for Better Schools (RBS). The project proposed to integrate advanced…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://files.eric.ed.gov/fulltext/EJ1066307.pdf','ERIC'); return false;" href="http://files.eric.ed.gov/fulltext/EJ1066307.pdf"><span>A Study of Effectiveness of Computer Assisted Instruction (<span class="hlt">CAI</span>) over Classroom Lecture (CRL) at ICS Level</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>Kaousar, Tayyeba; Choudhry, Bushra Naoreen; Gujjar, Aijaz Ahmed</p> <p>2008-01-01</p> <p>This study was aimed to evaluate the effectiveness of <span class="hlt">CAI</span> vs. classroom lecture for computer science at ICS level. The objectives were to compare the learning effects of two groups with classroom lecture and computer-assisted instruction studying the same curriculum and the effects of <span class="hlt">CAI</span> and CRL in terms of cognitive development. Hypotheses of…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20140012819','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20140012819"><span>Stable Magnesium Isotope Variation in Melilite Mantle of Allende Type B1 <span class="hlt">CAI</span> EK 459-5-1</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Kerekgyarto, A. G.; Jeffcoat, C. R.; Lapen, T. J.; Andreasen, R.; Righter, M.; Ross, D. K.</p> <p>2014-01-01</p> <p>Ca-Al-rich inclusions (<span class="hlt">CAIs</span>) are the earliest formed crystalline material in our solar system and they record early Solar System processes. Here we present petrographic and delta Mg-25 data of melilite mantles in a Type B1 <span class="hlt">CAI</span> that records early solar nebular processes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/9841405','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/9841405"><span>Oxygen reservoirs in the early solar nebula inferred from an Allende <span class="hlt">CAI</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Young, E D; Russell, S S</p> <p>1998-10-16</p> <p>Ultraviolet laser microprobe analyses of a calcium-aluminum-rich inclusion (<span class="hlt">CAI</span>) from the Allende meteorite suggest that a line with a slope of exactly 1.00 on a plot of delta (17)O against delta (18)O represents the primitive oxygen isotope reservoir of the early solar nebula. Most meteorites are enriched in (17)O and (18)O relative to this line, and their oxygen isotope ratios can be explained by mass fractionation or isotope exchange initiating from the primitive reservoir. These data establish a link between the oxygen isotopic composition of the abundant ordinary chondrites and the primitive (16)O-rich component of <span class="hlt">CAIs</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/9774267','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/9774267"><span>Oxygen reservoirs in the early solar nebula inferred from an allende <span class="hlt">CAI</span></span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Young; Russell</p> <p>1998-10-16</p> <p>Ultraviolet laser microprobe analyses of a calcium-aluminum-rich inclusion (<span class="hlt">CAI</span>) from the Allende meteorite suggest that a line with a slope of exactly 1.00 on a plot of delta17O against delta18O represents the primitive oxygen isotope reservoir of the early solar nebula. Most meteorites are enriched in 17O and 18O relative to this line, and their oxygen isotope ratios can be explained by mass fractionation or isotope exchange initiating from the primitive reservoir. These data establish a link between the oxygen isotopic composition of the abundant ordinary chondrites and the primitive 16O-rich component of <span class="hlt">CAIs</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19750022313','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19750022313"><span>Alternative communication network designs for an operational Plato 4 <span class="hlt">CAI</span> system</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Mobley, R. E., Jr.; Eastwood, L. F., Jr.</p> <p>1975-01-01</p> <p>The cost of alternative communications networks for the dissemination of PLATO IV computer-aided instruction (<span class="hlt">CAI</span>) was studied. Four communication techniques are compared: leased telephone lines, satellite communication, UHF TV, and low-power microwave radio. For each network design, costs per student contact hour are computed. These costs are derived as functions of student population density, a parameter which can be calculated from census data for one potential market for <span class="hlt">CAI</span>, the public primary and secondary schools. Calculating costs in this way allows one to determine which of the four communications alternatives can serve this market least expensively for any given area in the U.S. The analysis indicates that radio distribution techniques are cost optimum over a wide range of conditions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2008E%26PSL.272..353J','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2008E%26PSL.272..353J"><span>26Al- 26Mg and 207Pb- 206Pb systematics of Allende <span class="hlt">CAIs</span>: Canonical solar initial 26Al/ 27Al ratio reinstated</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Jacobsen, Benjamin; Yin, Qing-zhu; Moynier, Frederic; Amelin, Yuri; Krot, Alexander N.; Nagashima, Kazuhide; Hutcheon, Ian D.; Palme, Herbert</p> <p>2008-07-01</p> <p>The precise knowledge of the initial 26Al/ 27Al ratio [( 26Al/ 27Al) 0] is crucial if we are to use the very first solid objects formed in our Solar System, calcium-aluminum-rich inclusions (<span class="hlt">CAIs</span>) as the "time zero" age-anchor and guide future work with other short-lived radio-chronometers in the early Solar System, as well as determining the inventory of heat budgets from radioactivities for early planetary differentiation. New high-precision multi-collector inductively-coupled plasma mass spectrometry (MC-ICP-MS) measurements of 27Al/ 24Mg ratios and Mg-isotopic compositions of nine whole-rock <span class="hlt">CAIs</span> (six mineralogically characterized fragments and three micro-drilled inclusions) from the CV carbonaceous chondrite, Allende yield a well-defined 26Al- 26Mg fossil isochron with an ( 26Al/ 27Al) 0 of (5.23 ± 0.13) × 10 - 5 . Internal mineral isochrons obtained for three of these <span class="hlt">CAIs</span> ( A44A, AJEF, and A43) are consistent with the whole-rock <span class="hlt">CAI</span> isochron. The mineral isochron of AJEF with ( 26Al/ 27Al) 0 = (4.96 ± 0.25) × 10 - 5 , anchored to our precisely determined absolute 207Pb- 206Pb age of 4567.60 ± 0.36 Ma for the same mineral separates, reinstate the "canonical" ( 26Al/ 27Al) 0 of 5 × 10 - 5 for the early Solar System. The uncertainty in ( 26Al/ 27Al) 0 corresponds to a maximum time span of ± 20 Ka (thousand years), suggesting that the Allende <span class="hlt">CAI</span> formation events were culminated within this time span. Although all Allende <span class="hlt">CAIs</span> studied experienced multistage formation history, including melting and evaporation in the solar nebula and post-crystallization alteration likely on the asteroidal parent body, the 26Al- 26Mg and U-Pb-isotopic systematics of the mineral separates and bulk <span class="hlt">CAIs</span> behaved largely as closed-system since their formation. Our data do not support the "supra-canonical" 26Al/ 27Al ratio of individual minerals or their mixtures in CV <span class="hlt">CAIs</span>, suggesting that the supra-canonical 26Al/ 27Al ratio in the CV <span class="hlt">CAIs</span> may have resulted from post</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20140012818','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20140012818"><span>In Situ Trace Element Analysis of an Allende Type B1 <span class="hlt">CAI</span>: EK-459-5-1</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Jeffcoat, C. R.; Kerekgyarto, A.; Lapen, T. J.; Andreasen, R.; Righter, M.; Ross, D. K.</p> <p>2014-01-01</p> <p>Variations in refractory major and trace element composition of calcium, aluminum-rich inclusions (<span class="hlt">CAIs</span>) provide constraints on physical and chemical conditions and processes in the earliest stages of the Solar System. Previous work indicates that <span class="hlt">CAIs</span> have experienced complex histories involving, in many cases, multiple episodes of condensation, evaporation, and partial melting. We have analyzed major and trace element abundances in two core to rim transects of the melilite mantle as well as interior major phases of a Type B1 <span class="hlt">CAI</span> (EK-459-5-1) from Allende by electron probe micro-analyzer (EPMA) and laser ablation inductively coupled plasma mass spectrometry (LA-ICP-MS) to investigate the behavior of key trace elements with a primary focus on the REEs Tm and Yb.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JPhCS.983a2100Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JPhCS.983a2100Y"><span>The enhancement of students’ mathematical representation in junior high school using cognitive apprenticeship instruction (<span class="hlt">CAI</span>)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yusepa, B. G. P.; Kusumah, Y. S.; Kartasasmita, B. G.</p> <p>2018-03-01</p> <p>This study aims to get an in-depth understanding of the enhancement of students’ mathematical representation. This study is experimental research with pretest-posttest control group design. The subject of this study is the students’ of the eighth grade from junior high schools in Bandung: high-level and middle-level. In each school, two parallel groups were chosen as a control group and an experimental group. The experimental group was given cognitive apprenticeship instruction (<span class="hlt">CAI</span>) treatment while the control group was given conventional learning. The results show that the enhancement of students’ mathematical representation who obtained <span class="hlt">CAI</span> treatment was better than the conventional one, viewed which can be observed from the overall, mathematical prior knowledge (MPK), and school level. It can be concluded that <span class="hlt">CAI</span> can be used as a good alternative learning model to enhance students’ mathematical representation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19890006945','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19890006945"><span>Extending the granularity of representation and control for the MIL-STD <span class="hlt">CAIS</span> 1.0 node model</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Rogers, Kathy L.</p> <p>1986-01-01</p> <p>The Common APSE (Ada 1 Program Support Environment) Interface Set (<span class="hlt">CAIS</span>) (DoD85) node model provides an excellent baseline for interfaces in a single-host development environment. To encompass the entire spectrum of computing, however, the <span class="hlt">CAIS</span> model should be extended in four areas. It should provide the interface between the engineering workstation and the host system throughout the entire lifecycle of the system. It should provide a basis for communication and integration functions needed by distributed host environments. It should provide common interfaces for communications mechanisms to and among target processors. It should provide facilities for integration, validation, and verification of test beds extending to distributed systems on geographically separate processors with heterogeneous instruction set architectures (ISAS). Additions to the PROCESS NODE model to extend the <span class="hlt">CAIS</span> into these four areas are proposed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1994CPL...225...76C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1994CPL...225...76C"><span>Resonance-enhanced two-photon excitation of <span class="hlt">CaI</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Casero-Junquera, Elena; Lawruszczuk, Rafal; Rostas, Joëlle; Taieb, Guy</p> <p>1994-07-01</p> <p>Induced fluorescence following visible (620-655 nm) laser excitation of the <span class="hlt">CaI</span> radical has been detected not only in the same region (B, A-X transitions), but also in the UV (315-330 nm). The UV two-photon excitation spectrum consists of narrow bands appearing at laser frequencies located within certain bands of the Δ v = 1, 0 sequences of the B <span class="hlt">2</span>Σ +-X <span class="hlt">2</span>Σ + and A <span class="hlt">2</span>Π 1/<span class="hlt">2</span>-X <span class="hlt">2</span>Σ + systems. The main peaks are tentatively assigned to resonance-enhanced excitation of a single vibrational level of the lowest Rydberg D <span class="hlt">2</span>Σ + state from successive vibrational levels of the ground state. The excitation process is a one-color two-photon optical—optical-double-resonance via B <span class="hlt">2</span>Σ + and A <span class="hlt">2</span>Π 1/<span class="hlt">2</span> intermediate levels. This analysis is supported by the absorption spectrum observed long ago by Walters and Barratt. The absorption and laser excitation complementary data have been used to derive approximate molecular constants for the D state.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20100005633','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20100005633"><span>FIB-NanoSIMS-TEM Coordinated Study of a Wark-Lovering Rim in a Vigarano Type A <span class="hlt">CAI</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Cai, A.; Ito, M.; Keller, L. P.; Ross, D. K.; Nakamura-Messenger, K.</p> <p>2010-01-01</p> <p>Wark-Lovering (WL) rims are thin multi layered mineral sequences that surround most Ca, Al-rich inclusions (<span class="hlt">CAIs</span>). Unaltered WL rims are composed of the same primary high temperature minerals as <span class="hlt">CAIs</span>, such as melilite, spinel, pyroxene, hibonite, perovskite, anorthite and olivine. It is still unclear whether the rim minerals represent a different generation formed by a separate event from their associated <span class="hlt">CAIs</span> or are a byproduct of <span class="hlt">CAI</span> formation. Several models have been proposed for the origins of WL rims including condensation, flashheating, reaction of a <span class="hlt">CAI</span> with a Mg-Si-rich reservoir (nebular gas or solid); on the basis of mineralogy, abundances of trace elements, O and Mg isotopic studies. Detailed mineralogical characterizations of WL rims at micrometer to nanometer scales have been obtained by TEM observations, but so far no coordinated isotopic - mineralogical studies have been performed. Thus, we have applied an O isotopic imaging technique by NanoSIMS 50L to investigate heterogeneous distributions of O isotopic ratios in minerals within a cross section of a WL rim prepared using a focused ion beam (FIB) instrument. After the isotopic measurements, we determine the detailed mineralogy and microstructure of the same WL FIB section to gain insight into its petrogenesis. Here we present preliminary results from O isotopic and elemental maps by NanoSIMS and mineralogical analysis by FE-SEM of a FIB section of a WL rim in the Vigarano reduced CV3 chondrite.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018GeCoA.221..275D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018GeCoA.221..275D"><span>Titanium isotopes and rare earth patterns in <span class="hlt">CAIs</span>: Evidence for thermal processing and gas-dust decoupling in the protoplanetary disk</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Davis, Andrew M.; Zhang, Junjun; Greber, Nicolas D.; Hu, Jingya; Tissot, François L. H.; Dauphas, Nicolas</p> <p>2018-01-01</p> <p>Titanium isotopic compositions (mass-dependent fractionation and isotopic anomalies) were measured in 46 calcium-, aluminum-rich inclusions (<span class="hlt">CAIs</span>) from the Allende CV chondrite. After internal normalization to 49Ti/47Ti, we found that ε50Ti values are somewhat variable among <span class="hlt">CAIs</span>, and that ε46Ti is highly correlated with ε50Ti, with a best-fit slope of 0.162 ± 0.030 (95% confidence interval). The linear correlation between ε46Ti and ε50Ti extends the same correlation seen among bulk solar objects (slope 0.184 ± 0.007). This observation provides constraints on dynamic mixing of the solar disk and has implications for the nucleosynthetic origin of titanium isotopes, specifically on the possible contributions from various types of supernovae to the solar system. Titanium isotopic mass fractionation, expressed as δ‧49Ti, was measured by both sample-standard bracketing and double-spiking. Most <span class="hlt">CAIs</span> are isotopically unfractionated, within a 95% confidence interval of normal, but a few are significantly fractionated and the range δ‧49Ti is from ∼-4 to ∼+4. Rare earth element patterns were measured in 37 of the <span class="hlt">CAIs</span>. All <span class="hlt">CAIs</span> with significant titanium mass fractionation effects have group II and related REE patterns, implying kinetically controlled volatility fractionation during the formation of these <span class="hlt">CAIs</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_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('https://www.ncbi.nlm.nih.gov/pubmed/27722371','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27722371"><span>Thermal decomposition of sodium amide, <span class="hlt">Na</span>NH<span class="hlt">2</span>, and sodium amide hydroxide composites, <span class="hlt">Na</span>NH<span class="hlt">2</span>-<span class="hlt">Na</span>OH.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Jepsen, Lars H; Wang, Peikun; Wu, Guotao; Xiong, Zhitao; Besenbacher, Flemming; Chen, Ping; Jensen, Torben R</p> <p>2016-09-14</p> <p>Sodium amide, <span class="hlt">Na</span>NH <span class="hlt">2</span> , has recently been shown to be a useful catalyst to decompose NH 3 into H <span class="hlt">2</span> and N <span class="hlt">2</span> , however, sodium hydroxide is omnipresent and commercially available <span class="hlt">Na</span>NH <span class="hlt">2</span> usually contains impurities of <span class="hlt">Na</span>OH (<<span class="hlt">2</span>%). The thermal decomposition of <span class="hlt">Na</span>NH <span class="hlt">2</span> and <span class="hlt">Na</span>NH <span class="hlt">2</span> -<span class="hlt">Na</span>OH composites is systematically investigated and discussed. <span class="hlt">Na</span>NH <span class="hlt">2</span> is partially dissolved in <span class="hlt">Na</span>OH at T > 100 °C, forming a non-stoichiometric solid solution of <span class="hlt">Na</span>(OH) 1-x (NH <span class="hlt">2</span> ) x (0 < x < ∼0.30), which crystallizes in an orthorhombic unit cell with the space group P<span class="hlt">2</span> 1 <span class="hlt">2</span> 1 <span class="hlt">2</span> 1 determined by synchrotron powder X-ray diffraction. The composite x<span class="hlt">Na</span>NH <span class="hlt">2</span> -(1 - x)<span class="hlt">Na</span>OH (∼0.70 < x < 0.72) shows a lowered melting point, ∼160 °C, compared to 200 and 318 °C for neat <span class="hlt">Na</span>NH <span class="hlt">2</span> and <span class="hlt">Na</span>OH, respectively. We report that 0.36 mol of NH 3 per mol of <span class="hlt">Na</span>NH <span class="hlt">2</span> is released below 400 °C during heating in an argon atmosphere, initiated at its melting point, T = 200 °C, possibly due to the formation of the mixed sodium amide imide solid solution. Furthermore, <span class="hlt">Na</span>OH reacts with <span class="hlt">Na</span>NH <span class="hlt">2</span> at elevated temperatures and provides the release of additional NH 3 .</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016GeCoA.189...70K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016GeCoA.189...70K"><span>A link between oxygen, calcium and titanium isotopes in 26Al-poor hibonite-rich <span class="hlt">CAIs</span> from Murchison and implications for the heterogeneity of dust reservoirs in the solar nebula</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kööp, Levke; Davis, Andrew M.; Nakashima, Daisuke; Park, Changkun; Krot, Alexander N.; Nagashima, Kazuhide; Tenner, Travis J.; Heck, Philipp R.; Kita, Noriko T.</p> <p>2016-09-01</p> <p>PLACs (platy hibonite crystals) and related hibonite-rich calcium-, aluminum-rich inclusions (<span class="hlt">CAIs</span>; hereafter collectively referred to as PLAC-like <span class="hlt">CAIs</span>) have the largest nucleosynthetic isotope anomalies of all materials believed to have formed in the solar system. Most PLAC-like <span class="hlt">CAIs</span> have low inferred initial 26Al/27Al ratios and could have formed prior to injection or widespread distribution of 26Al in the solar nebula. In this study, we report 26Al-26Mg systematics combined with oxygen, calcium, and titanium isotopic compositions for a large number of newly separated PLAC-like <span class="hlt">CAIs</span> from the Murchison CM<span class="hlt">2</span> chondrite (32 <span class="hlt">CAIs</span> studied for oxygen, 26 of these also for 26Al-26Mg, calcium and titanium). Our results confirm (1) the large range of nucleosynthetic anomalies in 50Ti and 48Ca (our data range from -70‰ to +170‰ and -60‰ to +80‰, respectively), (<span class="hlt">2</span>) the substantial range of Δ17O values (-28‰ to -17‰, with Δ17O = δ17O - 0.52 × δ18O), and (3) general 26Al-depletion in PLAC-like <span class="hlt">CAIs</span>. The multielement approach reveals a relationship between Δ17O and the degree of variability in 50Ti and 48Ca: PLAC-like <span class="hlt">CAIs</span> with the highest Δ17O (∼-17‰) show large positive and negative 50Ti and 48Ca anomalies, while those with the lowest Δ17O (∼-28‰) have small to no anomalies in 50Ti and 48Ca. These observations could suggest a physical link between anomalous 48Ca and 50Ti carriers and an 16O-poor reservoir. We suggest that the solar nebula was isotopically heterogeneous shortly after collapse of the protosolar molecular cloud, and that the primordial dust reservoir, in which anomalous carrier phases were heterogeneously distributed, was 16O-poor (Δ17O ⩾ -17‰) relative to the primordial gaseous (CO + H<span class="hlt">2</span>O) reservoir (Δ17O < -35‰). However, other models such as CO self-shielding in the protoplanetary disk are also considered to explain the link between oxygen and calcium and titanium isotopes in PLAC-like <span class="hlt">CAIs</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/19929731','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/19929731"><span>Consumption of fa <span class="hlt">cai</span> Nostoc soup: a potential for BMAA exposure from Nostoc cyanobacteria in China?</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Roney, Britton R; Renhui, Li; Banack, Sandra Anne; Murch, Susan; Honegger, Rosmarie; Cox, Paul Alan</p> <p>2009-01-01</p> <p>Grown in arid regions of western China the cyanobacterium Nostoc flagelliforme--called fa <span class="hlt">cai</span> in Mandarin and fat choy in Cantonese--is wild-harvested and used to make soup consumed during New Year's celebrations. High prices, up to $125 USD/kg, led to overharvesting in Inner Mongolia, Ningxia, Gansu, Qinghai, and Xinjiang. Degradation of arid ecosystems, desertification, and conflicts between Nostoc harvesters and Mongol herdsmen concerned the Chinese environmental authorities, leading to a government ban of Nostoc commerce. This ban stimulated increased marketing of a substitute made from starch. We analysed samples purchased throughout China as well as in Chinese markets in the United States and the United Kingdom. Some were counterfeits consisting of dyed starch noodles. A few samples from California contained Nostoc flagelliforme but were adulterated with starch noodles. Other samples, including those from the United Kingdom, consisted of pure Nostoc flagelliforme. A recent survey of markets in Cheng Du showed no real Nostoc flagelliforme to be marketed. Real and artificial fa <span class="hlt">cai</span> differ in the presence of beta-N-methylamino-L-alanine (BMAA). Given its status as a high-priced luxury food, the government ban on collection and marketing, and the replacement of real fa <span class="hlt">cai</span> with starch substitutes consumed only on special occasions, it is anticipated that dietary exposure to BMAA from fa <span class="hlt">cai</span> will be reduced in the future in China.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2784433','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2784433"><span>Hunting and use of terrestrial fauna used by <span class="hlt">Cai</span>çaras from the Atlantic Forest coast (Brazil)</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p></p> <p>2009-01-01</p> <p>Background The Brazilian Atlantic Forest is considered one of the hotspots for conservation, comprising remnants of rain forest along the eastern Brazilian coast. Its native inhabitants in the Southeastern coast include the <span class="hlt">Cai</span>çaras (descendants from Amerindians and European colonizers), with a deep knowledge on the natural resources used for their livelihood. Methods We studied the use of the terrestrial fauna in three <span class="hlt">Cai</span>çara communities, through open-ended interviews with 116 native residents. Data were checked through systematic observations and collection of zoological material. Results The dependence on the terrestrial fauna by <span class="hlt">Cai</span>çaras is especially for food and medicine. The main species used are Didelphis spp., Dasyprocta azarae, Dasypus novemcinctus, and small birds (several species of Turdidae). Contrasting with a high dependency on terrestrial fauna resources by native Amazonians, the <span class="hlt">Cai</span>çaras do not show a constant dependency on these resources. Nevertheless, the occasional hunting of native animals represents a complimentary source of animal protein. Conclusion Indigenous or local knowledge on native resources is important in order to promote local development in a sustainable way, and can help to conserve biodiversity, particularly if the resource is sporadically used and not commercially exploited. PMID:19930595</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22080372-quantum-computational-universality-cai-miyake-duer-briegel-two-dimensional-quantum-state-from-affleck-kennedy-lieb-tasaki-quasichains','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22080372-quantum-computational-universality-cai-miyake-duer-briegel-two-dimensional-quantum-state-from-affleck-kennedy-lieb-tasaki-quasichains"><span>Quantum computational universality of the <span class="hlt">Cai</span>-Miyake-Duer-Briegel two-dimensional quantum state from Affleck-Kennedy-Lieb-Tasaki quasichains</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>Wei, Tzu-Chieh; C. N. Yang Institute for Theoretical Physics, State University of New York at Stony Brook, Stony Brook, New York 11794-3840; Raussendorf, Robert</p> <p>2011-10-15</p> <p>Universal quantum computation can be achieved by simply performing single-qubit measurements on a highly entangled resource state, such as cluster states. <span class="hlt">Cai</span>, Miyake, Duer, and Briegel recently constructed a ground state of a two-dimensional quantum magnet by combining multiple Affleck-Kennedy-Lieb-Tasaki quasichains of mixed spin-3/<span class="hlt">2</span> and spin-1/<span class="hlt">2</span> entities and by mapping pairs of neighboring spin-1/<span class="hlt">2</span> particles to individual spin-3/<span class="hlt">2</span> particles [Phys. Rev. A 82, 052309 (2010)]. They showed that this state enables universal quantum computation by single-spin measurements. Here, we give an alternative understanding of how this state gives rise to universal measurement-based quantum computation: by local operations, each quasichain canmore » be converted to a one-dimensional cluster state and entangling gates between two neighboring logical qubits can be implemented by single-spin measurements. We further argue that a two-dimensional cluster state can be distilled from the <span class="hlt">Cai</span>-Miyake-Duer-Briegel state.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/16093489','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/16093489"><span>Multimodality of Ca<span class="hlt">2</span>+ signaling in rat atrial myocytes.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Morad, Martin; Javaheri, Ashkan; Risius, Tim; Belmonte, Steve</p> <p>2005-06-01</p> <p>It has been suggested that the multiplicity of Ca(<span class="hlt">2</span>+) signaling pathways in atrial myocytes may contribute to the variability of its function. This article reports on a novel Ca(<span class="hlt">2</span>+) signaling cascade initiated by mechanical forces induced by "puffing" of solution onto the myocytes. <span class="hlt">Ca(i</span>) transients were measured in fura-<span class="hlt">2</span> acetoxymethyl (AM) loaded cells using alternating 340- and 410-nm excitation waves at 1.<span class="hlt">2</span> kHz. Pressurized puffs of bathing solutions, applied by an electronically controlled micro-barrel system, activated slowly (approximately 300 ms) developing <span class="hlt">Ca(i</span>) transients that lasted 1,693 +/- 68 ms at room temperature. Subsequent second and third puffs, applied at approximately 20 s intervals activated significantly smaller or no <span class="hlt">Ca(i</span>) transients. Puff-triggered <span class="hlt">Ca(i</span>) transients could be reactivated once again following caffeine (10 mM)-induced release of Ca(<span class="hlt">2</span>+) from sarcoplasmic reticulum (SR). Puff-triggered <span class="hlt">Ca(i</span>) transients were independent of [Ca(<span class="hlt">2</span>+)](o), and activation of voltage-gated Ca(<span class="hlt">2</span>+) or cationic stretch channels or influx of Ca(<span class="hlt">2</span>+) on <span class="hlt">Na</span>(+)/Ca(<span class="hlt">2</span>+)exchanger, because puffing solution containing no Ca(<span class="hlt">2</span>+), 10 microM diltiazem, 1 mM Cd(<span class="hlt">2</span>+), 5 mM Ni(<span class="hlt">2</span>+), or 100 microM Gd(3+) failed to suppress them. Puff-triggered <span class="hlt">Ca(i</span>) transients were enhanced in paced compared to quiescent myocytes. Electrically activated <span class="hlt">Ca(i</span>) transients triggered during the time course of puff-induced transients were unaltered, suggesting functionally separate Ca(<span class="hlt">2</span>+) pools. Contribution of inositol 1,4,5-triphosphate (IP(3))-gated or mitochondrial Ca(<span class="hlt">2</span>+) pools or modulation of SR stores by nitric oxide/nitric oxide synthase (NO/NOS) signaling were evaluated using 0.5 to 500 microM <span class="hlt">2</span>-aminoethoxydiphenyl borate (<span class="hlt">2</span>-APB) and 0.1 to 1 microM carbonylcyanide-p-trifluoromethoxyphenylhydrazone (FCCP), and 1 mM Nomega-Nitro-L-arginine methyl ester (L-NAME) and 7-nitroindizole, respectively. Only FCCP appeared to significantly suppress the puff-triggered <span class="hlt">Ca(i</span>) transients. It was</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://eric.ed.gov/?q=accounting+AND+fundamentals&pg=7&id=EJ550830','ERIC'); return false;" href="https://eric.ed.gov/?q=accounting+AND+fundamentals&pg=7&id=EJ550830"><span>Role of Computer Assisted Instruction (<span class="hlt">CAI</span>) in an Introductory Computer Concepts Course.</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>Skudrna, Vincent J.</p> <p>1997-01-01</p> <p>Discusses the role of computer assisted instruction (<span class="hlt">CAI</span>) in undergraduate education via a survey of related literature and specific applications. Describes an undergraduate computer concepts course and includes appendices of instructions, flowcharts, programs, sample student work in accounting, COBOL instructional model, decision logic in a…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26159472','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26159472"><span>Changes in flavour and microbial diversity during natural fermentation of suan-<span class="hlt">cai</span>, a traditional food made in Northeast China.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wu, Rina; Yu, Meiling; Liu, Xiaoyu; Meng, Lingshuai; Wang, Qianqian; Xue, Yating; Wu, Junrui; Yue, Xiqing</p> <p>2015-10-15</p> <p>We measured changes in the main physical and chemical properties, flavour compounds and microbial diversity in suan-<span class="hlt">cai</span> during natural fermentation. The results showed that the pH and concentration of soluble protein initially decreased but were then maintained at a stable level; the concentration of nitrite increased in the initial fermentation stage and after reaching a peak it decreased significantly to a low level by the end of fermentation. Suan-<span class="hlt">cai</span> was rich in 17 free amino acids. All of the free amino acids increased in concentration to different degrees, except histidine. Total free amino acids reached their highest levels in the mid-fermentation stage. The 17 volatile flavour components identified at the start of fermentation increased to 57 by the mid-fermentation stage; esters and aldehydes were in the greatest diversity and abundance, contributing most to the aroma of suan-<span class="hlt">cai</span>. Bacteria were more abundant and diverse than fungi in suan-<span class="hlt">cai</span>; 14 bacterial species were identified from the genera Leuconostoc, Bacillus, Pseudomonas and Lactobacillus. The predominant fungal species identified were Debaryomyces hansenii, Candida tropicalis and Penicillium expansum. Copyright © 2015 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011PhRvA..84d2333W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011PhRvA..84d2333W"><span>Quantum computational universality of the <span class="hlt">Cai</span>-Miyake-Dür-Briegel two-dimensional quantum state from Affleck-Kennedy-Lieb-Tasaki quasichains</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wei, Tzu-Chieh; Raussendorf, Robert; Kwek, Leong Chuan</p> <p>2011-10-01</p> <p>Universal quantum computation can be achieved by simply performing single-qubit measurements on a highly entangled resource state, such as cluster states. <span class="hlt">Cai</span>, Miyake, Dür, and Briegel recently constructed a ground state of a two-dimensional quantum magnet by combining multiple Affleck-Kennedy-Lieb-Tasaki quasichains of mixed spin-3/<span class="hlt">2</span> and spin-1/<span class="hlt">2</span> entities and by mapping pairs of neighboring spin-1/<span class="hlt">2</span> particles to individual spin-3/<span class="hlt">2</span> particles [Phys. Rev. APLRAAN1050-294710.1103/PhysRevA.82.052309 82, 052309 (2010)]. They showed that this state enables universal quantum computation by single-spin measurements. Here, we give an alternative understanding of how this state gives rise to universal measurement-based quantum computation: by local operations, each quasichain can be converted to a one-dimensional cluster state and entangling gates between two neighboring logical qubits can be implemented by single-spin measurements. We further argue that a two-dimensional cluster state can be distilled from the <span class="hlt">Cai</span>-Miyake-Dür-Briegel state.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://eric.ed.gov/?q=Body+AND+combat&pg=6&id=ED043228','ERIC'); return false;" href="https://eric.ed.gov/?q=Body+AND+combat&pg=6&id=ED043228"><span>Computer-Assisted Instruction in Engineering Dynamics. <span class="hlt">CAI</span>-Systems Memo Number 18.</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>Sheldon, John W.</p> <p></p> <p>A 90-minute computer-assisted instruction (<span class="hlt">CAI</span>) unit course supplemented by a 1-hour lecture on the dynamic nature of three-dimensional rotations and Euler angles was given to 29 undergraduate engineering students. The area of Euler angles was selected because it is essential to problem-working in three-dimensional rotations of a rigid body, yet…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4196991','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4196991"><span>Resveratrol Increases Nitric Oxide Production in the Rat Thick Ascending Limb via Ca<span class="hlt">2</span>+/Calmodulin</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Gonzalez-Vicente, Agustin; Cabral, Pablo D.; Garvin, Jeffrey L.</p> <p>2014-01-01</p> <p>The thick ascending limb of the loop of Henle reabsorbs 30% of the <span class="hlt">Na</span>Cl filtered through the glomerulus. Nitric oxide (NO) produced by NO synthase 3 (NOS3) inhibits <span class="hlt">Na</span>Cl absorption by this segment. Resveratrol, a polyphenol, has beneficial cardiovascular and renal effects, many of which are mediated by NO. Resveratrol increases intracellular Ca<span class="hlt">2</span>+ (<span class="hlt">Cai</span>) and AMP kinase (AMPK) and NAD-dependent deacetylase sirtuin1 (SIRT1) activities, all of which could activate NO production. We hypothesized that resveratrol stimulates NO production by thick ascending limbs via a Ca<span class="hlt">2</span>+/calmodulin-dependent mechanism. To test this, the effect of resveratrol on NO bioavailability was measured in thick ascending limb suspensions. <span class="hlt">Cai</span> was measured in single perfused thick ascending limbs. SIRT1 activity and expression were measured in thick ascending limb lysates. Resveratrol (100 µM) increased NO bioavailability in thick ascending limb suspensions by 1.3±0.<span class="hlt">2</span> AFU/mg/min (p<0.03). The NOS inhibitor L-NAME blunted resveratrol-stimulated NO bioavailability by 96±11% (p<0.03). The superoxide scavenger tempol had no effect. Resveratrol elevated <span class="hlt">Cai</span> from 48±7 to 135±24 nM (p<0.01) in single tubules. In Ca<span class="hlt">2</span>+-free media, the resveratrol-induced increase in NO was blunted by 60±20% (p<0.05) and the rise in <span class="hlt">Cai</span> reduced by 80%. Calmodulin inhibition prevented the resveratrol-induced increase in NO (p<0.002). AMPK inhibition had no effect. Resveratrol did not increase SIRT1 activity. We conclude that resveratrol increases NO production in thick ascending limbs via a Ca<span class="hlt">2</span>+/calmodulin dependent mechanism, and SIRT1 and AMPK do not participate. Resveratrol-stimulated NO production in thick ascending limbs may account for part of its beneficial effects. PMID:25314136</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25314136','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25314136"><span>Resveratrol increases nitric oxide production in the rat thick ascending limb via Ca<span class="hlt">2</span>+/calmodulin.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Gonzalez-Vicente, Agustin; Cabral, Pablo D; Garvin, Jeffrey L</p> <p>2014-01-01</p> <p>The thick ascending limb of the loop of Henle reabsorbs 30% of the <span class="hlt">Na</span>Cl filtered through the glomerulus. Nitric oxide (NO) produced by NO synthase 3 (NOS3) inhibits <span class="hlt">Na</span>Cl absorption by this segment. Resveratrol, a polyphenol, has beneficial cardiovascular and renal effects, many of which are mediated by NO. Resveratrol increases intracellular Ca<span class="hlt">2</span>+ (<span class="hlt">Cai</span>) and AMP kinase (AMPK) and NAD-dependent deacetylase sirtuin1 (SIRT1) activities, all of which could activate NO production. We hypothesized that resveratrol stimulates NO production by thick ascending limbs via a Ca<span class="hlt">2</span>+/calmodulin-dependent mechanism. To test this, the effect of resveratrol on NO bioavailability was measured in thick ascending limb suspensions. <span class="hlt">Cai</span> was measured in single perfused thick ascending limbs. SIRT1 activity and expression were measured in thick ascending limb lysates. Resveratrol (100 µM) increased NO bioavailability in thick ascending limb suspensions by 1.3±0.<span class="hlt">2</span> AFU/mg/min (p<0.03). The NOS inhibitor L-NAME blunted resveratrol-stimulated NO bioavailability by 96±11% (p<0.03). The superoxide scavenger tempol had no effect. Resveratrol elevated <span class="hlt">Cai</span> from 48±7 to 135±24 nM (p<0.01) in single tubules. In Ca<span class="hlt">2</span>+-free media, the resveratrol-induced increase in NO was blunted by 60±20% (p<0.05) and the rise in <span class="hlt">Cai</span> reduced by 80%. Calmodulin inhibition prevented the resveratrol-induced increase in NO (p<0.002). AMPK inhibition had no effect. Resveratrol did not increase SIRT1 activity. We conclude that resveratrol increases NO production in thick ascending limbs via a Ca<span class="hlt">2</span>+/calmodulin dependent mechanism, and SIRT1 and AMPK do not participate. Resveratrol-stimulated NO production in thick ascending limbs may account for part of its beneficial effects.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20160002232','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20160002232"><span>New Petrology, Mineral Chemistry and Stable MG Isotope Compositions of an Allende <span class="hlt">CAI</span>: EK-459-7-<span class="hlt">2</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Jeffcoat, C. R.; Kerekgyarto, A. G.; Lapen, T. J.; Righter, M.; Simon, J. I.; Ross, D. K.</p> <p>2016-01-01</p> <p>Calcium-aluminum-rich inclusions (<span class="hlt">CAIs</span>) are the key to understanding physical and chemical conditions in the nascent solar nebula. These inclusions have the oldest radiometric ages of solar system materials and are composed of phases that are predicted to condense early from a gas of solar composition. Thus, their chemistry and textures record conditions and processes in the earliest stages of development of the solar nebula. Type B inclusions are typically larger and more coarse grained than other types with substantial evidence that many of them were at least partially molten. Type B inclusions are further subdivided into Type B1 (possess thick melilite mantle) and Type B<span class="hlt">2</span> (lack melilite mantle). Despite being extensively studied, the origin of the melilite mantles of Type B1 inclusions remains uncertain. We present petrologic and chemical data for a Type B inclusion, EK-459-7-<span class="hlt">2</span>, that bears features found in both Type B1 and B<span class="hlt">2</span> inclusions and likely represents an intermediate between the two types. Detailed studies of more of these intermediate objects may help to constrain models for Type B1 rim formation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://eric.ed.gov/?q=histology&pg=5&id=EJ635928','ERIC'); return false;" href="https://eric.ed.gov/?q=histology&pg=5&id=EJ635928"><span>Web Pages: An Effective Method of Providing <span class="hlt">CAI</span> Resource Material in Histology.</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>McLean, Michelle</p> <p>2001-01-01</p> <p>Presents research that introduces computer-aided instruction (<span class="hlt">CAI</span>) resource material as an integral part of the second-year histology course at the University of Natal Medical School. Describes the ease with which this software can be developed, using limited resources and available skills, while providing students with valuable learning…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2005LPI....36.1525Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2005LPI....36.1525Y"><span>Supra-Canonical Initial 26Al/27Al Indicate a 105 Year Residence Time for <span class="hlt">CAIs</span> in the Solar Proto-Planetary Disk</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Young, E. D.; Simon, J. I.; Galy, A.; Russell, S. S.; Tonui, E. K.; Lovera, O.</p> <p>2005-03-01</p> <p>We present new UV laser ablation and acid digestion MC-ICPMS analyses of 8 <span class="hlt">CAIs</span> showing that there was more 26Al in the early solar system than previously thought, and that the canonical initial 26Al/27Al represents a ~300,000 yr residence time for <span class="hlt">CAIs</span> in the protoplanetary disk.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017LPICo1987.6381D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017LPICo1987.6381D"><span>The Range of Initial 10Be/9Be Ratios in the Early Solar System: A Re-Assessment Based on Analyses of New <span class="hlt">CAIs</span> and Melilite Composition Glass Standards</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Dunham, E.; Wadhwa, M.; Liu, M.-C.</p> <p>2017-07-01</p> <p>We report a more accurate range of initial 10Be/9Be in <span class="hlt">CAIs</span> including FUN <span class="hlt">CAI</span> CMS-1 from Allende (CV3) and a new <span class="hlt">CAI</span> from NWA 5508 (CV3) using melilite composition glass standards; we suggest 10Be is largely produced by irradiation in the nebula.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://files.eric.ed.gov/fulltext/ED074769.pdf','ERIC'); return false;" href="http://files.eric.ed.gov/fulltext/ED074769.pdf"><span>Evaluation of a Text Compression Algorithm Against Computer-Aided Instruction (<span class="hlt">CAI</span>) Material.</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>Knight, Joseph M., Jr.</p> <p></p> <p>This report describes the initial evaluation of a text compression algorithm against computer assisted instruction (<span class="hlt">CAI</span>) material. A review of some concepts related to statistical text compression is followed by a detailed description of a practical text compression algorithm. A simulation of the algorithm was programed and used to obtain…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29407387','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29407387"><span>Bacterial and fungal microbiota of spontaneously fermented Chinese products, Rubing milk cake and Yan-<span class="hlt">cai</span> vegetable pickles.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Liu, Xin; Kuda, Takashi; Takahashi, Hajime; Kimura, Bon</p> <p>2018-06-01</p> <p>The Rubing milk cake from Yunnan and the Yan-<span class="hlt">cai</span> vegetable pickles from Guangdong are traditional spontaneously fermented foods in China. We evaluated the microbial properties of these products with the analysis of their bacterial and fungal microbiota using classical culture-dependent and culture-independent methods, including a 16S rDNA gene (V4) and an internal transcribed spacer (ITS) region pyrosequencing method with MiSeq system. The viable lactic acid bacteria (LAB) count was 8 and 6 log colony-forming units (CFU)/g in Rubing and Yan-<span class="hlt">cai</span> samples, respectively. The yeast count was approximately 100-1000 times less than the LAB count in most samples, except one Yan-<span class="hlt">cai</span> sample. In addition, the gram-negative rod count in half of the samples was similar to the LAB count. Pyrosequencing results revealed the high abundance (10%-20%) of gram-negative Pseudomonas spp. and Enterobacteriaceae in these samples. These results suggest that some of these traditional foods are undesirable as ready-to-eat (RTE) foods, even when these are typical lactic acid fermented foods. Copyright © 2017 Elsevier Ltd. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://eric.ed.gov/?q=Should+AND+programming+AND+taught&pg=2&id=ED295668','ERIC'); return false;" href="https://eric.ed.gov/?q=Should+AND+programming+AND+taught&pg=2&id=ED295668"><span>A CBI Model for the Design of <span class="hlt">CAI</span> Software by Teachers/Nonprogrammers.</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>Tessmer, Martin; Jonassen, David H.</p> <p></p> <p>This paper describes a design model presented in workbook form which is intended to facilitate computer-assisted instruction (<span class="hlt">CAI</span>) software design by teachers who do not have programming experience. Presentation of the model is preceded by a number of assumptions that underlie the instructional content and methods of the textbook. It is argued…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.A21A2152H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.A21A2152H"><span>Adaptation of an aerosol retrieval algorithm using multi-wavelength and multi-pixel information of satellites (MWPM) to GOSAT/TANSO-<span class="hlt">CAI</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hashimoto, M.; Takenaka, H.; Higurashi, A.; Nakajima, T.</p> <p>2017-12-01</p> <p>Aerosol in the atmosphere is an important constituent for determining the earth's radiation budget, so the accurate aerosol retrievals from satellite is useful. We have developed a satellite remote sensing algorithm to retrieve the aerosol optical properties using multi-wavelength and multi-pixel information of satellite imagers (MWPM). The method simultaneously derives aerosol optical properties, such as aerosol optical thickness (AOT), single scattering albedo (SSA) and aerosol size information, by using spatial difference of wavelegths (multi-wavelength) and surface reflectances (multi-pixel). The method is useful for aerosol retrieval over spatially heterogeneous surface like an urban region. In this algorithm, the inversion method is a combination of an optimal method and smoothing constraint for the state vector. Furthermore, this method has been combined with the direct radiation transfer calculation (RTM) numerically solved by each iteration step of the non-linear inverse problem, without using look up table (LUT) with several constraints. However, it takes too much computation time. To accelerate the calculation time, we replaced the RTM with an accelerated RTM solver learned by neural network-based method, EXAM (Takenaka et al., 2011), using Rster code. And then, the calculation time was shorternd to about one thouthandth. We applyed MWPM combined with EXAM to GOSAT/TANSO-<span class="hlt">CAI</span> (Cloud and Aerosol Imager). <span class="hlt">CAI</span> is a supplement sensor of TANSO-FTS, dedicated to measure cloud and aerosol properties. <span class="hlt">CAI</span> has four bands, 380, 674, 870 and 1600 nm, and observes in 500 meters resolution for band1, band<span class="hlt">2</span> and band3, and 1.5 km for band4. Retrieved parameters are aerosol optical properties, such as aerosol optical thickness (AOT) of fine and coarse mode particles at a wavelenth of 500nm, a volume soot fraction in fine mode particles, and ground surface albedo of each observed wavelength by combining a minimum reflectance method and Fukuda et al. (2013). We will show</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/1998PhDT.......222B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1998PhDT.......222B"><span>An investigative study into the effectiveness of using computer-aided instruction (<span class="hlt">CAI</span>) as a laboratory component of college-level biology: A case study</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Barrett, Joan Beverly</p> <p></p> <p>Community colleges serve the most diverse student populations in higher education. They consist of non-traditional, part-time, older, intermittent, and mobile students of different races, ethnic backgrounds, language preferences, physical and mental abilities, and learning style preferences. Students who are academically challenged may have diverse learning characteristics that are not compatible with the more traditional approaches to the delivery of instruction. With this need come new ways of solving the dilemma, such as Computer-aided Instruction (<span class="hlt">CAI</span>). This case study investigated the use of <span class="hlt">CAI</span> as a laboratory component of college-level biology in a small, rural community college setting. The intent was to begin to fill a void that seems to exist in the literature regarding the role of the faculty in the development and use of <span class="hlt">CAI</span>. In particular, the investigator was seeking to understand the practice and its effectiveness, especially in helping the under prepared student. The case study approach was chosen to examine a specific phenomenon within a single institution. Ethnographic techniques, such as interviewing, documentary analysis, life's experiences, and participant observations were used to collect data about the phenomena being studied. Results showed that the faculty was primarily self-motivated and self-taught in their use of <span class="hlt">CAI</span> as a teaching and learning tool. The importance of faculty leadership and collegiality was evident. Findings showed the faculty confident that expectations of helping students who have difficulties with mathematical concepts have been met and that <span class="hlt">CAI</span> is becoming the most valuable of learning tools. In a traditional college classroom, or practice, time is the constant (semesters) and competence is the variable. In the <span class="hlt">CAI</span> laboratory time became the variable and competence the constant. The use of <span class="hlt">CAI</span> also eliminated hazardous chemicals that were routinely used in the more traditional lab. Outcomes showed that annual savings</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20070009991','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20070009991"><span>Isotopic Measurements in <span class="hlt">CAIs</span> with the Nanosims: Implications to the understanding of the Formation process of Ca, Al-Rich Inclusions</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Ito, M.; Messenger, S.; Walker, Robert M.</p> <p>2007-01-01</p> <p>Ca, Al-rich Inclusions (<span class="hlt">CAIs</span>) preserve evidence of thermal events that they experienced during their formation in the early solar system. Most <span class="hlt">CAIs</span> from CV and CO chondrites are characterized by large variations in O-isotopic compositions of primary minerals, with spinel, hibonite, and pyroxene being more O-16-rich than melilite and anorthite, with delta 17, O-18 = approx. -40%o (DELTA O-17 = delta O-17 - 0.52 x delta O-18 = approx. - 20%o ). These anomalous compositions cannot be accounted for by standard mass dependent fractionation and diffusive process of those minerals. It requires the presence of an anomalous oxygen reservoir of nucleosynthetic origin or mass independent fractionations before the formation of <span class="hlt">CAIs</span> in the early solar system. The CAMECA NanoSIMS is a new generation ion microprobe that offers high sensitivity isotopic measurements with sub 100 nm spatial resolution. The NanoSIMS has significantly improved abilities in the study of presolar grains in various kind of meteorites and the decay products of extinct nuclides in ancient solar system matter. This instrument promises significant improvements over other conventional ion probes in the precision isotopic characterization of sub-micron scales. We report the results of our first O isotopic measurements of various <span class="hlt">CAI</span> minerals from EK1-6-3 and 7R19-1(a) utilizing the JSC NanoSIMS 50L ion microprobe. We evaluate the measurement conditions, the instrumental mass fractionation factor (IMF) for O isotopic measurement and the accuracy of the isotopic ratio through the analysis of a San Carlos olivine standard and <span class="hlt">CAI</span> sample of 7R19-1(a).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JSSCh.258..416S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JSSCh.258..416S"><span><span class="hlt">Na</span>7 [Fe<span class="hlt">2</span>S6 ] , <span class="hlt">Na</span><span class="hlt">2</span> [FeS<span class="hlt">2</span> ] and <span class="hlt">Na</span><span class="hlt">2</span> [FeSe<span class="hlt">2</span> ] : New 'reduced' sodium chalcogenido ferrates</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Stüble, Pirmin; Peschke, Simon; Johrendt, Dirk; Röhr, Caroline</p> <p>2018-02-01</p> <p>Three new 'reduced' FeII containing sodium chalcogenido ferrates were obtained applying a reductive synthetic route. The mixed-valent sulfido ferrate <span class="hlt">Na</span>7 [Fe<span class="hlt">2</span>S6 ] , which forms bar-shaped crystals with metallic greenish luster, was synthesized in pure phase from natural pyrite and elemental sodium at a maximum temperature of 800 °C. Its centrosymmetric triclinic structure (SG P 1 bar , a = 764.15(<span class="hlt">2</span>), b = 1153.70(<span class="hlt">2</span>), c = 1272.58(3) pm, α = 62.3325 (7) , β = 72.8345 (8) , γ = 84.6394 (8) ° , Z = 3, R1 = 0.0185) exhibits two crystallographically different [Fe<span class="hlt">2</span>S6 ] 7 - dimers of edge-sharing [FeS4 ] tetrahedra, with somewhat larger Fe-S distances than in the fully oxidized FeIII dimers of e.g. <span class="hlt">Na</span>6 [Fe<span class="hlt">2</span>III S6 ] . In contrast to the localized AFM ordered pure di-ferrates(III), the Curie-Weiss behavior of the magnetic susceptibility proves the rarely observed valence-delocalized S = 9/<span class="hlt">2</span> state of the mixed-valent FeIII /FeII dimer. The nearly spin-only value of the magnetic moment combined with the chemical bonding not generally differing from that in pure ferrates(II) and (III), provides a striking argument, that the reduction of the local Fe spin moments observed in all condensed sulfido ferrate moieties is connected with the AFM spin ordering. The two isotypic ferrates(II) <span class="hlt">Na</span><span class="hlt">2</span> [FeS<span class="hlt">2</span> ] and <span class="hlt">Na</span><span class="hlt">2</span> [FeSe<span class="hlt">2</span> ] with chain-like structural units (SG Ibam, a = 643.54(8)/ 660.81(1), b = 1140.<span class="hlt">2(2)/1190.30(2</span>) c = 562.90(6)/585.59(1) pm, Z = 4, R1 = 0.0372/0.0466) crystallize in the K<span class="hlt">2</span> [ZnO<span class="hlt">2</span> ] -type structure. Although representing merely further members of the common series of chalcogenido metallates(II) <span class="hlt">Na</span><span class="hlt">2</span> [MIIQ<span class="hlt">2</span> ] , these two new phases, together with <span class="hlt">Na</span>6 [FeS4 ] and Li<span class="hlt">2</span> [FeS<span class="hlt">2</span> ] , are the only examples of pure FeII alkali chalcogenido ferrates. The new compounds allow for a general comparison of di- and chain ferrates(II) and (III) and mixed-valent analogs concerning the electronic and magnetic properties (including Heisenberg super-exchange and double-exchange interactions</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2006IJCEM...7...41C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2006IJCEM...7...41C"><span>Numerical Investigation Into Effect of Fuel Injection Timing on <span class="hlt">CAI</span>/HCCI Combustion in a Four-Stroke GDI Engine</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Cao, Li; Zhao, Hua; Jiang, Xi; Kalian, Navin</p> <p>2006-02-01</p> <p>The Controlled Auto-Ignition (<span class="hlt">CAI</span>) combustion, also known as Homogeneous Charge Compression Ignition (HCCI), was achieved by trapping residuals with early exhaust valve closure in conjunction with direct injection. Multi-cycle 3D engine simulations have been carried out for parametric study on four different injection timings in order to better understand the effects of injection timings on in-cylinder mixing and <span class="hlt">CAI</span> combustion. The full engine cycle simulation including complete gas exchange and combustion processes was carried out over several cycles in order to obtain the stable cycle for analysis. The combustion models used in the present study are the Shell auto-ignition model and the characteristic-time combustion model, which were modified to take the high level of EGR into consideration. A liquid sheet breakup spray model was used for the droplet breakup processes. The analyses show that the injection timing plays an important role in affecting the in-cylinder air/fuel mixing and mixture temperature, which in turn affects the <span class="hlt">CAI</span> combustion and engine performance.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20150018570','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20150018570"><span>Characterization of Meteorites by Focused Ion Beam Sectioning: Recent Applications to <span class="hlt">CAIs</span> and Primitive Meteorite Matrices</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Christoffersen, Roy; Keller, Lindsay P.; Han, Jangmi; Rahman, Zia; Berger, Eve L.</p> <p>2015-01-01</p> <p>Focused ion beam (FIB) sectioning has revolutionized preparation of meteorite samples for characterization by analytical transmission electron microscopy (TEM) and other techniques. Although FIB is not "non-destructive" in the purest sense, each extracted section amounts to no more than nanograms (approximately 500 cubic microns) removed intact from locations precisely controlled by SEM imaging and analysis. Physical alteration of surrounding material by ion damage, fracture or sputter contamination effects is localized to within a few micrometers around the lift-out point. This leaves adjacent material intact for coordinate geochemical analysis by SIMS, microdrill extraction/TIMS and other techniques. After lift out, FIB sections can be quantitatively analyzed by electron microprobe prior to final thinning, synchrotron x-ray techniques, and by the full range of state-of-the-art analytical field-emission scanning transmission electron microscope (FE-STEM) techniques once thinning is complete. Multiple meteorite studies supported by FIB/FE-STEM are currently underway at NASA-JSC, including coordinated analysis of refractory phase assemblages in <span class="hlt">CAIs</span> and fine-grained matrices in carbonaceous chondrites. FIB sectioning of <span class="hlt">CAIs</span> has uncovered epitaxial and other overgrowth relations between corundum-hibonite-spinel consistent with hibonite preceding corundum and/or spinel in non-equilibrium condensation sequences at combinations of higher gas pressures, dust-gas enrichments or significant nebular transport. For all of these cases, the ability of FIB to allow for coordination with spatially-associated isotopic data by SIMS provides immense value for constraining the formation scenarios of the particular <span class="hlt">CAI</span> assemblage. For carbonaceous chondrites matrix material, FIB has allowed us to obtain intact continuous sections of the immediate outer surface of Murchison (CM<span class="hlt">2</span>) after it has been experimentally ion processed to simulate solar wind space weathering. The surface</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://eric.ed.gov/?q=software+AND+component+AND+documentation&pg=2&id=ED070262','ERIC'); return false;" href="https://eric.ed.gov/?q=software+AND+component+AND+documentation&pg=2&id=ED070262"><span>Everything You Always Wanted to Know About <span class="hlt">CAI</span> But Were Afraid To Ask.</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>Luskin, Bernard J.; And Others</p> <p></p> <p>A comprehensive summary of significant developments related to the integration of the computer in all levels of instruction, this book identifies, classifies, and examines obstacles to computer-assisted instruction (<span class="hlt">CAI</span>), their scope and possible resolutions. Some 75 experts were surveyed and their opinions statistically analyzed in regard to 23…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010JChPh.132x4305Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010JChPh.132x4305Z"><span>Long range intermolecular interactions between the alkali diatomics <span class="hlt">Na</span><span class="hlt">2</span>, K<span class="hlt">2</span>, and <span class="hlt">Na</span>K</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zemke, Warren T.; Byrd, Jason N.; Michels, H. Harvey; Montgomery, John A.; Stwalley, William C.</p> <p>2010-06-01</p> <p>Long range interactions between the ground state alkali diatomics <span class="hlt">Na</span><span class="hlt">2</span>-<span class="hlt">Na</span><span class="hlt">2</span>, K<span class="hlt">2</span>-K<span class="hlt">2</span>, <span class="hlt">Na</span><span class="hlt">2</span>-K<span class="hlt">2</span>, and <span class="hlt">NaK-Na</span>K are examined. Interaction energies are first determined from ab initio calculations at the coupled-cluster with singles, doubles, and perturbative triples [CCSD(T)] level of theory, including counterpoise corrections. Long range energies calculated from diatomic molecular properties (polarizabilities and dipole and quadrupole moments) are then compared with the ab initio energies. A simple asymptotic model potential ELR=Eelec+Edisp+Eind is shown to accurately represent the intermolecular interactions for these systems at long range.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70023467','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70023467"><span>Thermal maturity patterns in New York State using <span class="hlt">CAI</span> and %Ro</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Weary, D.J.; Ryder, R.T.; Nyahay, R.E.</p> <p>2001-01-01</p> <p>New conodont alteration index (<span class="hlt">CAI</span>) and vitrinite reflectance (%Ro) data collected from drill holes in the Appalachian basin of New York State allow refinement of thermal maturity maps for Ordovician and Devonian rocks. <span class="hlt">CAI</span> isotherms on the new maps show a pattern that approximates that published by Harris et al. (1978) in eastern and western New York, but it differs in central New York, where the isotherms are shifted markedly westward by more than 100 km and are more tightly grouped. This close grouping of isograds reflects a steeper thermal gradient than previously noted by Harris et al. (1978) and agrees closely with the abrupt west-to-east increase in thermal maturity across New York noted by Johnsson (1986). These data show, in concordance with previous studies, that thermal maturity levels in these rocks are higher than can be explained by simple burial heating beneath the present thickness of overburden. The Ordovician and Devonian rocks of the Appalachian Basin in New York must have been buried by very thick post-Devonian sediments (4-6 km suggested by Sarwar and Friedman 1995) or were exposed to a higher-than-normal geothermal flux caused by crustal extension, or a combination of the two.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29616791','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29616791"><span>On the Stability of <span class="hlt">Na</span>O<span class="hlt">2</span> in <span class="hlt">Na</span>-O<span class="hlt">2</span> Batteries.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Liu, Chenjuan; Carboni, Marco; Brant, William R; Pan, Ruijun; Hedman, Jonas; Zhu, Jiefang; Gustafsson, Torbjörn; Younesi, Reza</p> <p>2018-04-25</p> <p><span class="hlt">Na</span>-O <span class="hlt">2</span> batteries are regarded as promising candidates for energy storage. They have higher energy efficiency, rate capability, and chemical reversibility than Li-O <span class="hlt">2</span> batteries; in addition, sodium is cheaper and more abundant compared to lithium. However, inconsistent observations and instability of discharge products have inhibited the understanding of the working mechanism of this technology. In this work, we have investigated a number of factors that influence the stability of the discharge products. By means of in operando powder X-ray diffraction study, the influence of oxygen, sodium anode, salt, solvent, and carbon cathode were investigated. The <span class="hlt">Na</span> metal anode and an ether-based solvent are the main factors that lead to the instability and decomposition of <span class="hlt">Na</span>O <span class="hlt">2</span> in the cell environment. This fundamental insight brings new information on the working mechanism of <span class="hlt">Na</span>-O <span class="hlt">2</span> batteries.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.dtic.mil/docs/citations/ADA139278','DTIC-ST'); return false;" href="http://www.dtic.mil/docs/citations/ADA139278"><span>A <span class="hlt">CAI</span> (Computer-Assisted Instruction) Course on Constructing PLANIT lessons: Development, Content, and Evaluation</span></a></p> <p><a target="_blank" href="http://www.dtic.mil/">DTIC Science & Technology</a></p> <p></p> <p>1980-06-01</p> <p>courseware package on how to program lessons for an automated system. Since PLANIT (Programming Language for Interactive Teaching) is the student/author...assisted instruction (<span class="hlt">CAI</span>), how to program PLANIT lessons, and to evaluate the effectiveness of the package for select Army users. The resultant courseware</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/20590191','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/20590191"><span>Long range intermolecular interactions between the alkali diatomics <span class="hlt">Na</span>(<span class="hlt">2</span>), K(<span class="hlt">2</span>), and <span class="hlt">Na</span>K.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zemke, Warren T; Byrd, Jason N; Michels, H Harvey; Montgomery, John A; Stwalley, William C</p> <p>2010-06-28</p> <p>Long range interactions between the ground state alkali diatomics <span class="hlt">Na</span>(<span class="hlt">2</span>)-<span class="hlt">Na</span>(<span class="hlt">2</span>), K(<span class="hlt">2</span>)-K(<span class="hlt">2</span>), <span class="hlt">Na</span>(<span class="hlt">2</span>)-K(<span class="hlt">2</span>), and <span class="hlt">NaK-Na</span>K are examined. Interaction energies are first determined from ab initio calculations at the coupled-cluster with singles, doubles, and perturbative triples [CCSD(T)] level of theory, including counterpoise corrections. Long range energies calculated from diatomic molecular properties (polarizabilities and dipole and quadrupole moments) are then compared with the ab initio energies. A simple asymptotic model potential E(LR)=E(elec)+E(disp)+E(ind) is shown to accurately represent the intermolecular interactions for these systems at long range.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://files.eric.ed.gov/fulltext/ED503459.pdf','ERIC'); return false;" href="http://files.eric.ed.gov/fulltext/ED503459.pdf"><span>A Comparative Study to Evaluate the Effectiveness of Computer Assisted Instruction (<span class="hlt">CAI</span>) versus Class Room Lecture (RL) for Computer Science at ICS Level</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>Kausar, Tayyaba; Choudhry, Bushra Naoreen; Gujjar, Aijaz Ahmed</p> <p>2008-01-01</p> <p>This study was aimed to evaluate the effectiveness of <span class="hlt">CAI</span> vs. classroom lecture for computer science at ICS level. The objectives were to compare the learning effects of two groups with class room lecture and computer assisted instruction studying the same curriculum and the effects of <span class="hlt">CAI</span> and CRL in terms of cognitive development. Hypothesis of…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://files.eric.ed.gov/fulltext/EJ1102933.pdf','ERIC'); return false;" href="http://files.eric.ed.gov/fulltext/EJ1102933.pdf"><span>A Comparative Study to Evaluate the Effectiveness of Computer Assisted Instruction (<span class="hlt">CAI</span>) versus Class Room Lecture (CRL) for Computer Science at ICS Level</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>Kausar, Tayyaba; Choudhry, Bushra Naoreen; Gujjar, Aijaz Ahmed</p> <p>2008-01-01</p> <p>This study was aimed to evaluate the effectiveness of <span class="hlt">CAI</span> vs. classroom lecture for computer science at ICS level. The objectives were to compare the learning effects of two groups with class room lecture and computer assisted instruction studying the same curriculum and the effects of <span class="hlt">CAI</span> and CRL in terms of cognitive development. Hypothesis of…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017GeCoA.207....1T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017GeCoA.207....1T"><span>In situ isotopic studies of the U-depleted Allende <span class="hlt">CAI</span> Curious Marie: Pre-accretionary alteration and the co-existence of 26Al and 36Cl in the early solar nebula</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Tang, Haolan; Liu, Ming-Chang; McKeegan, Kevin D.; Tissot, Francois L. H.; Dauphas, Nicolas</p> <p>2017-06-01</p> <p>The isotopic composition of oxygen as well as 26Al-26Mg and 36Cl-36S systematics were studied in Curious Marie, an aqueously altered Allende <span class="hlt">CAI</span> characterized by a Group II REE pattern and a large 235U excess produced by the decay of short-lived 247Cm. Oxygen isotopic compositions in the secondary minerals of Curious Marie follow a mass-dependent fractionation line with a relatively homogenous depletion in 16O (Δ17O of -8‰) compared to unaltered minerals of <span class="hlt">CAI</span> components. Both Mg and S show large excesses of radiogenic isotopes (26Mg∗ and 36S∗) that are uniformly distributed within the <span class="hlt">CAI</span>, independent of parent/daughter ratio. A model initial 26Al/27Al ratio [(6.<span class="hlt">2</span> ± 0.9) × 10-5], calculated using the bulk Al/Mg ratio and the uniform δ26Mg∗ ∼ +43‰, is similar to the canonical initial solar system value within error. The exceptionally high bulk Al/Mg ratio of this <span class="hlt">CAI</span> (∼95) compared to other inclusions is presumably due to Mg mobilization by fluids. Therefore, the model initial 26Al/27Al ratio of this <span class="hlt">CAI</span> implies not only the early condensation of the <span class="hlt">CAI</span> precursor but also that aqueous alteration occurred early, when 26Al was still at or near the canonical value. This alteration event is most likely responsible for the U depletion in Curious Marie and occurred at most 50 kyr after <span class="hlt">CAI</span> formation, leading to a revised estimate of the early solar system 247Cm/235U ratio of (5.6 ± 0.3) × 10-5. The Mg isotopic composition in Curious Marie was subsequently homogenized by closed-system thermal processing without contamination by chondritic Mg. The large, homogeneous 36S excesses (Δ36S∗ ∼ +97‰) detected in the secondary phases of Curious Marie are attributed to 36Cl decay (t1/<span class="hlt">2</span> = 0.3 Myr) that was introduced by Cl-rich fluids during the aqueous alteration event that led to sodalite formation. A model 36Cl/35Cl ratio of (<span class="hlt">2</span>.3 ± 0.6) × 10-5 is calculated at the time of aqueous alteration, translating into an initial 36Cl/35Cl ratio of ∼1.7-3 </p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29310275','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29310275"><span><span class="hlt">CaI</span> and SrI molecules for iodine determination by high-resolution continuum source graphite furnace molecular absorption spectrometry: Greener molecules for practical application.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zanatta, Melina Borges Teixeira; Nakadi, Flávio Venâncio; da Veiga, Márcia Andreia Mesquita Silva</p> <p>2018-03-01</p> <p>A new method to determine iodine in drug samples by high-resolution continuum source graphite furnace molecular absorption spectrometry (HR-CS GF MAS) has been developed. The method measures the molecular absorption of a diatomic molecule, <span class="hlt">CaI</span> or SrI (less toxic molecule-forming reagents), at 638.904 or 677.692nm, respectively, and uses a mixture containing 5μg of Pd and 0.5μg of Mg as chemical modifier. The method employs pyrolysis temperatures of 1000 and 800°C and vaporization temperatures of 2300 and 2400°C for <span class="hlt">CaI</span> and SrI, respectively. The optimized amounts of Ca and Sr as molecule-forming reagents are 100 and 150µg, respectively. On the basis of interference studies, even small chlorine concentrations reduce <span class="hlt">CaI</span> and SrI absorbance significantly. The developed method was used to analyze different commercial drug samples, namely thyroid hormone pills with three different iodine amounts (15.88, 31.77, and 47.66µg) and one liquid drug with 1% m v -1 active iodine in their compositions. The results agreed with the values informed by the manufacturers (95% confidence level) regardless of whether <span class="hlt">CaI</span> or SrI was determined. Therefore, the developed method is useful for iodine determination on the basis of <span class="hlt">CaI</span> or SrI molecular absorption. Copyright © 2017 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://eric.ed.gov/?q=artificial+AND+intelligence+AND+business&pg=3&id=EJ613257','ERIC'); return false;" href="https://eric.ed.gov/?q=artificial+AND+intelligence+AND+business&pg=3&id=EJ613257"><span>A Cross-National <span class="hlt">CAI</span> Tool To Support Learning Operations Decision-Making and Market Analysis.</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>Mockler, Robert J.; Afanasiev, Mikhail Y.; Dologite, Dorothy G.</p> <p>1999-01-01</p> <p>Describes bicultural (United States and Russia) development of a computer-aided instruction (<span class="hlt">CAI</span>) tool to learn management decision-making using information systems technologies. The program has been used with undergraduate and graduate students in both countries; it integrates free and controlled market concepts and combines traditional computer…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4376416','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4376416"><span><span class="hlt">Na</span>+/Ca<span class="hlt">2</span>+ exchange and <span class="hlt">Na</span>+/K+-ATPase in the heart</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Shattock, Michael J; Ottolia, Michela; Bers, Donald M; Blaustein, Mordecai P; Boguslavskyi, Andrii; Bossuyt, Julie; Bridge, John H B; Chen-Izu, Ye; Clancy, Colleen E; Edwards, Andrew; Goldhaber, Joshua; Kaplan, Jack; Lingrel, Jerry B; Pavlovic, Davor; Philipson, Kenneth; Sipido, Karin R; Xie, Zi-Jian</p> <p>2015-01-01</p> <p>This paper is the third in a series of reviews published in this issue resulting from the University of California Davis Cardiovascular Symposium 2014: Systems approach to understanding cardiac excitation–contraction coupling and arrhythmias: <span class="hlt">Na</span>+ channel and <span class="hlt">Na</span>+ transport. The goal of the symposium was to bring together experts in the field to discuss points of consensus and controversy on the topic of sodium in the heart. The present review focuses on cardiac <span class="hlt">Na</span>+/Ca<span class="hlt">2</span>+ exchange (NCX) and <span class="hlt">Na</span>+/K+-ATPase (NKA). While the relevance of Ca<span class="hlt">2</span>+ homeostasis in cardiac function has been extensively investigated, the role of <span class="hlt">Na</span>+ regulation in shaping heart function is often overlooked. Small changes in the cytoplasmic <span class="hlt">Na</span>+ content have multiple effects on the heart by influencing intracellular Ca<span class="hlt">2</span>+ and pH levels thereby modulating heart contractility. Therefore it is essential for heart cells to maintain <span class="hlt">Na</span>+ homeostasis. Among the proteins that accomplish this task are the <span class="hlt">Na</span>+/Ca<span class="hlt">2</span>+ exchanger (NCX) and the <span class="hlt">Na</span>+/K+ pump (NKA). By transporting three <span class="hlt">Na</span>+ ions into the cytoplasm in exchange for one Ca<span class="hlt">2</span>+ moved out, NCX is one of the main <span class="hlt">Na</span>+ influx mechanisms in cardiomyocytes. Acting in the opposite direction, NKA moves <span class="hlt">Na</span>+ ions from the cytoplasm to the extracellular space against their gradient by utilizing the energy released from ATP hydrolysis. A fine balance between these two processes controls the net amount of intracellular <span class="hlt">Na</span>+ and aberrations in either of these two systems can have a large impact on cardiac contractility. Due to the relevant role of these two proteins in <span class="hlt">Na</span>+ homeostasis, the emphasis of this review is on recent developments regarding the cardiac <span class="hlt">Na</span>+/Ca<span class="hlt">2</span>+ exchanger (NCX1) and <span class="hlt">Na</span>+/K+ pump and the controversies that still persist in the field. PMID:25772291</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20160002651','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20160002651"><span>Calcium and Titanium Isotope Fractionation in <span class="hlt">CAIS</span>: Tracers of Condensation and Inheritance in the Early Solar Protoplanetary Disk</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Simon, J. I.; Jordan, M. K.; Tappa, M. J.; Kohl, I. E.; Young, E. D.</p> <p>2016-01-01</p> <p>The chemical and isotopic compositions of calcium-aluminum-rich inclusions (<span class="hlt">CAIs</span>) can be used to understand the conditions present in the protoplantary disk where they formed. The isotopic compositions of these early-formed nebular materials are largely controlled by chemical volatility. The isotopic effects of evaporation/sublimation, which are well explained by both theory and experimental work, lead to enrichments of the heavy isotopes that are often exhibited by the moderately refractory elements Mg and Si. Less well understood are the isotopic effects of condensation, which limits our ability to determine whether a <span class="hlt">CAI</span> is a primary condensate and/or retains any evidence of its primordial formation history.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22141459','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22141459"><span>Developing the Coach Analysis and Intervention System (<span class="hlt">CAIS</span>): establishing validity and reliability of a computerised systematic observation instrument.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Cushion, Christopher; Harvey, Stephen; Muir, Bob; Nelson, Lee</p> <p>2012-01-01</p> <p>We outline the evolution of a computerised systematic observation tool and describe the process for establishing the validity and reliability of this new instrument. The Coach Analysis and Interventions System (<span class="hlt">CAIS</span>) has 23 primary behaviours related to physical behaviour, feedback/reinforcement, instruction, verbal/non-verbal, questioning and management. The instrument also analyses secondary coach behaviour related to performance states, recipient, timing, content and questioning/silence. The <span class="hlt">CAIS</span> is a multi-dimensional and multi-level mechanism able to provide detailed and contextualised data about specific coaching behaviours occurring in complex and nuanced coaching interventions and environments that can be applied to both practice sessions and competition.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2232891','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2232891"><span>Electrogenic <span class="hlt">Na</span>+/Ca<span class="hlt">2</span>+ Exchange</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Danaceau, Jonathan P.; Lucero, Mary T.</p> <p>2000-01-01</p> <p>Olfactory receptor neurons (ORNs) from the squid, Lolliguncula brevis, respond to the odors l-glutamate or dopamine with increases in internal Ca<span class="hlt">2</span>+ concentrations ([Ca<span class="hlt">2</span>+]i). To directly asses the effects of increasing [Ca<span class="hlt">2</span>+]i in perforated-patched squid ORNs, we applied 10 mM caffeine to release Ca<span class="hlt">2</span>+ from internal stores. We observed an inward current response to caffeine. Monovalent cation replacement of <span class="hlt">Na</span>+ from the external bath solution completely and selectively inhibited the caffeine-induced response, and ruled out the possibility of a Ca<span class="hlt">2</span>+-dependent nonselective cation current. The strict dependence on internal Ca<span class="hlt">2</span>+ and external <span class="hlt">Na</span>+ indicated that the inward current was due to an electrogenic <span class="hlt">Na</span>+/Ca<span class="hlt">2</span>+ exchanger. Block of the caffeine-induced current by an inhibitor of <span class="hlt">Na</span>+/Ca<span class="hlt">2</span>+ exchange (50–100 μM <span class="hlt">2</span>′,4′-dichlorobenzamil) and reversibility of the exchanger current, further confirmed its presence. We tested whether <span class="hlt">Na</span>+/Ca<span class="hlt">2</span>+ exchange contributed to odor responses by applying the aquatic odor l-glutamate in the presence and absence of <span class="hlt">2</span>′,4′-dichlorobenzamil. We found that electrogenic <span class="hlt">Na</span>+/Ca<span class="hlt">2</span>+ exchange was responsible for ∼26% of the total current associated with glutamate-induced odor responses. Although <span class="hlt">Na</span>+/Ca<span class="hlt">2</span>+ exchangers are known to be present in ORNs from numerous species, this is the first work to demonstrate amplifying contributions of the exchanger current to odor transduction. PMID:10828249</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li class="active"><span>8</span></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_8 --> <div id="page_9" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li class="active"><span>9</span></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="161"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/21390725-experimental-study-fuel-injection-strategies-cai-gasoline-engine','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/21390725-experimental-study-fuel-injection-strategies-cai-gasoline-engine"><span>An experimental study of fuel injection strategies in <span class="hlt">CAI</span> gasoline engine</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>Hunicz, J.; Kordos, P.</p> <p>2011-01-15</p> <p>Combustion of gasoline in a direct injection controlled auto-ignition (<span class="hlt">CAI</span>) single-cylinder research engine was studied. <span class="hlt">CAI</span> operation was achieved with the use of the negative valve overlap (NVO) technique and internal exhaust gas re-circulation (EGR). Experiments were performed at single injection and split injection, where some amount of fuel was injected close to top dead centre (TDC) during NVO interval, and the second injection was applied with variable timing. Additionally, combustion at variable fuel-rail pressure was examined. Investigation showed that at fuel injection into recompressed exhaust fuel reforming took place. This process was identified via an analysis of the exhaust-fuelmore » mixture composition after NVO interval. It was found that at single fuel injection in NVO phase, its advance determined the heat release rate and auto-ignition timing, and had a strong influence on NO{sub X} emission. However, a delay of single injection to intake stroke resulted in deterioration of cycle-to-cycle variability. Application of split injection showed benefits of this strategy versus single injection. Examinations of different fuel mass split ratios and variable second injection timing resulted in further optimisation of mixture formation. At equal share of the fuel mass injected in the first injection during NVO and in the second injection at the beginning of compression, the lowest emission level and cyclic variability improvement were observed. (author)« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://files.eric.ed.gov/fulltext/ED432263.pdf','ERIC'); return false;" href="http://files.eric.ed.gov/fulltext/ED432263.pdf"><span>Using Pre-test/Post-test Data To Evaluate the Effectiveness of Computer Aided Instruction (A Study of <span class="hlt">CAI</span> and Its Use with Developmental Reading Students).</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>Lansford, Carl E.</p> <p></p> <p>As computer aided instruction (<span class="hlt">CAI</span>) and distance learning become more popular, a model for easily evaluating these teaching methods must be developed, one which will enable replication of the study each year. This paper discusses the results of a study using existing dependent and independent variables to evaluate <span class="hlt">CAI</span> for developmental reading…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1992MTB....23..833K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1992MTB....23..833K"><span><span class="hlt">Na</span><span class="hlt">2</span>O-Al<span class="hlt">2</span>O3 system: Activity of <span class="hlt">Na</span><span class="hlt">2</span>O in (α + β)- and (β + β)-alumina</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kale, G. M.</p> <p>1992-12-01</p> <p>The activity of <span class="hlt">Na</span><span class="hlt">2</span>O in a biphasic mixture of (α + β)-alumina has been measured in the temperature range of 700 to 1100 K using the solid-state galvanic cell: 11663_2007_Article_BF02656462_TeX<span class="hlt">2</span>GIFE1.gif _{(1:1)}^{Pt,CO_<span class="hlt">2</span> + O_<span class="hlt">2</span> /<span class="hlt">Na</span>_<span class="hlt">2</span> CO_3 /(α + β ) - alumin a//(Y_<span class="hlt">2</span> O_3 )ZrO_<span class="hlt">2</span> //In + In_<span class="hlt">2</span> O_3 ,Ta,Pt} Similarly, the activity of <span class="hlt">Na</span><span class="hlt">2</span>O in a (β + β’’)-alumina two-phase mixture has been measured between 700 and 1100 K employing the galvanic cell: 11663_2007_Article_BF02656462_TeX<span class="hlt">2</span>GIFE<span class="hlt">2</span>.gif _{(1:1)}^{Pt,CO_<span class="hlt">2</span> + O_<span class="hlt">2</span> /<span class="hlt">Na</span>_<span class="hlt">2</span> CO_3 /(β + β ) - alumin a//(Y_<span class="hlt">2</span> O_3 )ZrO_<span class="hlt">2</span> //In + In_<span class="hlt">2</span> O_3 ,Ta,Pt} The reversible electromotive force (emf ) of both the cells was found to vary linearly with temperature over the entire temperature range of measurement. From the measured reversible emf and auxiliary thermodynamic data for In<span class="hlt">2</span>O<span class="hlt">2</span>, <span class="hlt">Na</span><span class="hlt">2</span>O, CO<span class="hlt">2</span> and <span class="hlt">Na</span><span class="hlt">2</span>CO3 reported in the literature, the temperature dependence of the logarithm of activity of <span class="hlt">Na</span><span class="hlt">2</span>O in (α + β)-alumina is obtained: 11663_2007_Article_BF02656462_TeX<span class="hlt">2</span>GIFE3.gif log α _{<span class="hlt">Na</span>_<span class="hlt">2</span> O} (α + β ) = 1.85 - 14,750/T(K)( ± 0.015)(700 ≤slant T ≤slant 1100) For (β + β'’)-alumina, 11663_2007_Article_BF02656462_TeX<span class="hlt">2</span>GIFE4.gif log α _{<span class="hlt">Na</span>_<span class="hlt">2</span> O} (β + β ) = 3.9 - 13,000/T(K)( ± 0.015)(700 ≤slant T ≤slant 1100)</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/11884373','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/11884373"><span>L-type Ca(<span class="hlt">2</span>+) currents overlapping threshold <span class="hlt">Na</span>(+) currents: could they be responsible for the "slip-mode" phenomenon in cardiac myocytes?</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Piacentino, Valentino; Gaughan, John P; Houser, Steven R</p> <p>2002-03-08</p> <p>Phosphorylation of <span class="hlt">Na</span> channels has been suggested to increase their Ca permeability. Termed "slip-mode conductance" (SMC), this hypothesis predicts that Ca influx via protein kinase A (PKA)-modified <span class="hlt">Na</span> channels can induce sarcoplasmic reticulum (SR) Ca release. We tested this hypothesis by determining if SR Ca release is graded with I(<span class="hlt">Na</span>) in the presence of activated PKA (with Isoproterenol, ISO). V(m), I(m), and [<span class="hlt">Ca](i</span>) were measured in feline (n=26) and failing human (n=19) ventricular myocytes. Voltage steps from -70 through -40 mV were used to grade I(<span class="hlt">Na</span>). <span class="hlt">Na</span> channel antagonists (tetrodotoxin), L-type Ca channel (I(Ca,L)) antagonists (nifedipine, cadmium, verapamil), and agonists (Bay K 8644, FPL 64176) were used to separate SMC from I(Ca,L). In the absence of ISO, I(<span class="hlt">Na</span>) was associated with SR Ca release in human but not feline myocytes. After ISO, graded I(<span class="hlt">Na</span>) was associated with small amounts of SR Ca release in feline myocytes and the magnitude of release increased in human myocytes. I(<span class="hlt">Na</span>)-related SR Ca release was insensitive to tetrodotoxin (n=10) but was blocked by nifedipine (n=10) and cadmium (n=3). SR Ca release was induced over the same voltage range in the absence of ISO with Bay K 8644 and FPL 64176 (n=9). Positive voltage steps (to 0 mV) to fully activate <span class="hlt">Na</span> channels (SMC) in the presence of ISO and Verapamil only caused SR Ca release when block of I(Ca,L) was incomplete. We conclude that PKA-mediated increases in I(Ca,L) and SR Ca loading can reproduce many of the experimental features of SMC.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26522496','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26522496"><span>The contribution of the androgen receptor (AR) in human spatial learning and memory: A study in women with complete androgen insensitivity syndrome (<span class="hlt">CAIS</span>).</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Mueller, S C; Verwilst, T; Van Branteghem, A; T'Sjoen, G; Cools, M</p> <p>2016-02-01</p> <p>Few studies have examined the impact of androgen insensitivity on human spatial learning and memory. In the present study, we tested 11 women with complete androgen insensitivity syndrome (<span class="hlt">CAIS</span>), a rare genetic disorder characterized by complete absence of AR activity, and compared their performance against 20 comparison males and 19 comparison females on a virtual analog of the Morris Water Maze task. The results replicated a main sex effect showing that men relative to women were faster in finding the hidden platform and had reduced heading error. Furthermore, findings indicated that mean performance of women with <span class="hlt">CAIS</span> was between control women and control men, though the differences were not statistically significant. Effect size estimates (and corresponding confidence intervals) of spatial learning trials showed little difference between women with <span class="hlt">CAIS</span> and control women but <span class="hlt">CAIS</span> women differed from men, but not women, on two variables, latency to find the platform and first-move latency. No differences between groups were present during visible platform trials or the probe trial, a measure of spatial memory. Moreover, groups also did not differ on estimates of IQ and variability of performance. The findings are discussed in relation to androgen insensitivity in human spatial learning and memory. Copyright © 2015 Elsevier Inc. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/FR-2012-02-17/pdf/2012-3839.pdf','FEDREG'); return false;" href="https://www.gpo.gov/fdsys/pkg/FR-2012-02-17/pdf/2012-3839.pdf"><span>77 FR 9625 - Presentation of Final Conventional Conformance Test Criteria and Common Air Interface (<span class="hlt">CAI</span>...</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collection.action?collectionCode=FR">Federal Register 2010, 2011, 2012, 2013, 2014</a></p> <p></p> <p>2012-02-17</p> <p>... Tests for Inclusion in the Program AGENCY: National Institute of Standards and Technology (NIST... meeting is to present the final requirements for <span class="hlt">CAI</span> conventional conformance tests for inclusion in the... suitability for inclusion in the P25 CAP is below: Conformance tests should limit devices in the test...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JPS...377..121S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JPS...377..121S"><span>Controlled phase stability of highly <span class="hlt">Na</span>-active triclinic structure in nanoscale high-voltage <span class="hlt">Na</span><span class="hlt">2-2</span>xCo1+xP<span class="hlt">2</span>O7 cathode for <span class="hlt">Na</span>-ion batteries</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Song, Hee Jo; Kim, Jae-Chan; Dar, Mushtaq Ahmad; Kim, Dong-Wan</p> <p>2018-02-01</p> <p>With the increasing demand for high energy density in energy-storage systems, a high-voltage cathode is essential in rechargeable Li-ion and <span class="hlt">Na</span>-ion batteries. The operating voltage of a triclinic-polymorph <span class="hlt">Na</span><span class="hlt">2</span>CoP<span class="hlt">2</span>O7, also known as the rose form, is above 4.0 V (vs. <span class="hlt">Na/Na</span>+), which is relatively high compared to that of other cathode materials. Thus, it can be employed as a potential high-voltage cathode material in <span class="hlt">Na</span>-ion batteries. However, it is difficult to synthesize a pure rose phase because of its low phase stability, thus limiting its use in high-voltage applications. Herein, compositional-engineered, rose-phase <span class="hlt">Na</span><span class="hlt">2-2</span>xCo1+xP<span class="hlt">2</span>O7/C (x = 0, 0.1 and 0.<span class="hlt">2</span>) nanopowder are prepared using a wet-chemical method. The <span class="hlt">Na</span><span class="hlt">2-2</span>xCo1+xP<span class="hlt">2</span>O7/C cathode shows high electrochemical reactivity with <span class="hlt">Na</span> ions at 4.0 V, delivering high capacity and high energy density.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29734805','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29734805"><span>High-Performance <span class="hlt">Na</span>-O<span class="hlt">2</span> Batteries Enabled by Oriented <span class="hlt">Na</span>O<span class="hlt">2</span> Nanowires as Discharge Products.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Khajehbashi, S Mohammad B; Xu, Lin; Zhang, Guobin; Tan, Shuangshuang; Zhao, Yan; Wang, Lai-Sen; Li, Jiantao; Luo, Wen; Peng, Dong-Liang; Mai, Liqiang</p> <p>2018-06-13</p> <p><span class="hlt">Na</span>-O <span class="hlt">2</span> batteries are emerging rechargeable batteries due to their high theoretical energy density and abundant resources, but they suffer from sluggish kinetics due to the formation of large-size discharge products with cubic or irregular particle shapes. Here, we report the unique growth of discharge products of <span class="hlt">Na</span>O <span class="hlt">2</span> nanowires inside <span class="hlt">Na</span>-O <span class="hlt">2</span> batteries that significantly boosts the performance of <span class="hlt">Na</span>-O <span class="hlt">2</span> batteries. For this purpose, a high-spin Co 3 O 4 electrocatalyst was synthesized via the high-temperature oxidation of pure cobalt nanoparticles in an external magnetic field. The discharge products of <span class="hlt">Na</span>O <span class="hlt">2</span> nanowires are 10-20 nm in diameter and ∼10 μm in length, characteristics that provide facile pathways for electron and ion transfer. With these nanowires, <span class="hlt">Na</span>-O <span class="hlt">2</span> batteries have surpassed 400 cycles with a fixed capacity of 1000 mA h g -1 , an ultra-low over-potential of ∼60 mV during charging, and near-zero over-potential during discharging. This strategy not only provides a unique way to control the morphology of discharge products to achieve high-performance <span class="hlt">Na</span>-O <span class="hlt">2</span> batteries but also opens up the opportunity to explore growing nanowires in novel conditions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011JMoSt1006..547N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011JMoSt1006..547N"><span>RNA adducts with <span class="hlt">Na</span> <span class="hlt">2</span>SeO 4 and <span class="hlt">Na</span> <span class="hlt">2</span>SeO 3 - Stability and structural features</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Nafisi, Shohreh; Manouchehri, Firouzeh; Montazeri, Maryam</p> <p>2011-12-01</p> <p>Selenium compounds are widely available in dietary supplements and have been extensively studied for their antioxidant and anticancer properties. Low blood Se levels were found to be associated with an increased incidence and mortality from various types of cancers. Although many in vivo and clinical trials have been conducted using these compounds, their biochemical and chemical mechanisms of efficacy are the focus of much current research. This study was designed to examine the interaction of <span class="hlt">Na</span> <span class="hlt">2</span>SeO 4 and <span class="hlt">Na</span> <span class="hlt">2</span>SeO 3 with RNA in aqueous solution at physiological conditions, using a constant RNA concentration (6.25 mM) and various sodium selenate and sodium selenite/polynucleotide (phosphate) ratios of 1/80, 1/40, 1/20, 1/10, 1/5, 1/<span class="hlt">2</span> and 1/1. Fourier transform infrared, UV-Visible spectroscopic methods were used to determine the drug binding modes, the binding constants, and the stability of <span class="hlt">Na</span> <span class="hlt">2</span>SeO 4 and <span class="hlt">Na</span> <span class="hlt">2</span>SeO 3-RNA complexes in aqueous solution. Spectroscopic evidence showed that <span class="hlt">Na</span> <span class="hlt">2</span>SeO 4 and <span class="hlt">Na</span> <span class="hlt">2</span>SeO 3 bind to the major and minor grooves of RNA ( via G, A and U bases) with some degree of the Se-phosphate (PO <span class="hlt">2</span>) interaction for both compounds with overall binding constants of K(<span class="hlt">Na</span> <span class="hlt">2</span>SeO 4-RNA) = 8.34 × 10 3 and K(<span class="hlt">Na</span> <span class="hlt">2</span>SeO 3-RNA) = 4.57 × 10 3 M -1. The order of selenium salts-biopolymer stability was <span class="hlt">Na</span> <span class="hlt">2</span>SeO 4-RNA > <span class="hlt">Na</span> <span class="hlt">2</span>SeO 3-RNA. RNA aggregations occurred at higher selenium concentrations. No biopolymer conformational changes were observed upon <span class="hlt">Na</span> <span class="hlt">2</span>SeO 4 and <span class="hlt">Na</span> <span class="hlt">2</span>SeO 3 interactions, while RNA remains in the A-family structure.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29433594','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29433594"><span>The effectiveness of self-regulation in limiting the advertising of unhealthy foods and beverages on children's preferred websites in Canada.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Potvin Kent, Monique; Pauzé, Elise</p> <p>2018-06-01</p> <p>To assess the effectiveness of the self-regulatory Canadian Children's Food and Beverage Advertising Initiative (<span class="hlt">CAI</span>) in limiting advertising of unhealthy foods and beverages on children's preferred websites in Canada.Design/Setting/SubjectsSyndicated Internet advertising exposure data were used to identify the ten most popular websites for children (aged <span class="hlt">2</span>-11 years) and determine the frequency of food/beverage banner and pop-up ads on these websites from June 2015 to May 2016. Nutrition information for advertised products was collected and their nutrient content per 100 g was calculated. Nutritional quality of all food/beverage ads was assessed using the Pan American Health Organization (PAHO) and UK Nutrient Profile Models (NPM). Nutritional quality of <span class="hlt">CAI</span> and non-<span class="hlt">CAI</span> company ads was compared using χ <span class="hlt">2</span> analyses and independent t tests. About 54 million food/beverage ads were viewed on children's preferred websites from June 2015 to May 2016. Most (93·4 %) product ads were categorized as excessive in fat, <span class="hlt">Na</span> or free sugars as per the PAHO NPM and 73·8 % were deemed less healthy according to the UK NPM. <span class="hlt">CAI</span>-company ads were <span class="hlt">2</span>·<span class="hlt">2</span> times more likely (OR; 99 % CI) to be excessive in at least one nutrient (<span class="hlt">2</span>·<span class="hlt">2</span>; <span class="hlt">2</span>·1, <span class="hlt">2</span>·<span class="hlt">2</span>, P<0·001) and <span class="hlt">2</span>·5 times more likely to be deemed less healthy (<span class="hlt">2</span>·5; <span class="hlt">2</span>·5, <span class="hlt">2</span>·5, P<0·001) than non-<span class="hlt">CAI</span> ads. On average, <span class="hlt">CAI</span>-company product ads also contained (mean difference; 99 % CI) more energy (141; 141·1, 141·4 kcal, P<0·001, r=0·55), sugar (18·<span class="hlt">2</span>; 18·<span class="hlt">2</span>, 18·<span class="hlt">2</span> g, P<0·001, r=0·68) and <span class="hlt">Na</span> (70·0; 69·7, 70·0 mg, P<0·001, r=0·23) per 100 g serving than non-<span class="hlt">CAI</span> ads. The <span class="hlt">CAI</span> is not limiting unhealthy food and beverage advertising on children's preferred websites in Canada. Mandatory regulations are needed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19930068398&hterms=Israel+humanity&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3DIsrael%2Bhumanity','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19930068398&hterms=Israel+humanity&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3DIsrael%2Bhumanity"><span>Secondary processing of chondrules and refractory inclusions (<span class="hlt">CAIs</span>) by gasdynamic heating</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Podolak, M.; Prialnik, D.; Bunch, T. E.; Cassen, P.; Reynolds, R.</p> <p>1993-01-01</p> <p>Results of calculations performed to determine the conditions necessary for producing the opaque rims on chondrules and <span class="hlt">CAI</span> rims by high-speed entry into the transient atmosphere of an accreting meteorite parent body are presented. The sensitivity of these results to variations in critical parameters is investigated. The range of entry velocities which can produce such rims is shown to depend on the size, melting temperature, and thermal conductivity of the particles. For particles greater than <span class="hlt">2</span> mm in radius, with thermal conductivities of 20,000 ergs/sm s K or lower, entry velocities of about 3 km/s suffice. For particle sizes less than 1 mm in radius, the range of encounter velocities that can produce rims is narrow or vanishing, regardless of the thermal conductivity, unless the melting temperature in the outer part of the chondrule has been reduced by compositional heterogeneity.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22584108-neutron-diffraction-studies-na-ion-battery-electrode-materials-nacocr-sub-po-sub-sub-nanicr-sub-po-sub-sub-na-sub-ni-sub-cr-po-sub-sub','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22584108-neutron-diffraction-studies-na-ion-battery-electrode-materials-nacocr-sub-po-sub-sub-nanicr-sub-po-sub-sub-na-sub-ni-sub-cr-po-sub-sub"><span>Neutron diffraction studies of the <span class="hlt">Na</span>-ion battery electrode materials <span class="hlt">Na</span>CoCr{sub <span class="hlt">2</span>}(PO{sub 4}){sub 3}, <span class="hlt">Na</span>NiCr{sub <span class="hlt">2</span>}(PO{sub 4}){sub 3}, and <span class="hlt">Na</span>{sub <span class="hlt">2</span>}Ni{sub <span class="hlt">2</span>}Cr(PO{sub 4}){sub 3}</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>Yahia, H. Ben; Essehli, R., E-mail: ressehli@qf.org.qa; Avdeev, M.</p> <p></p> <p>The new compounds <span class="hlt">Na</span>CoCr{sub <span class="hlt">2</span>}(PO{sub 4}){sub 3}, <span class="hlt">Na</span>NiCr{sub <span class="hlt">2</span>}(PO{sub 4}){sub 3}, and <span class="hlt">Na</span>{sub <span class="hlt">2</span>}Ni{sub <span class="hlt">2</span>}Cr(PO{sub 4}){sub 3} were synthesized by sol-gel method and their crystal structures were determined by using neutron powder diffraction data. These compounds were characterized by galvanometric cycling and cyclic voltammetry. <span class="hlt">Na</span>CoCr{sub <span class="hlt">2</span>}(PO{sub 4}){sub 3}, <span class="hlt">Na</span>NiCr{sub <span class="hlt">2</span>}(PO{sub 4}){sub 3}, and <span class="hlt">Na</span>{sub <span class="hlt">2</span>}Ni{sub <span class="hlt">2</span>}Cr(PO{sub 4}){sub 3} crystallize with a stuffed α-CrPO{sub 4}-type structure. The structure consists of a 3D-framework made of octahedra and tetrahedra that are sharing corners and/or edges generating channels along [100] and [010], in which the sodium atoms are located. Of significance, in the structuresmore » of <span class="hlt">Na</span>NiCr{sub <span class="hlt">2</span>}(PO{sub 4}){sub 3}, and <span class="hlt">Na</span>{sub <span class="hlt">2</span>}Ni{sub <span class="hlt">2</span>}Cr(PO{sub 4}){sub 3} a statistical disorder Ni{sup <span class="hlt">2</span>+}/Cr{sup 3+} was observed on both the 8g and 4a atomic positions, whereas in <span class="hlt">Na</span>CoCr{sub <span class="hlt">2</span>}(PO{sub 4}){sub 3} the statistical disorder Co{sup <span class="hlt">2</span>+}/Cr{sup 3+} was only observed on the 8g atomic position. When tested as negative electrode materials, <span class="hlt">Na</span>CoCr{sub <span class="hlt">2</span>}(PO{sub 4}){sub 3}, <span class="hlt">Na</span>NiCr{sub <span class="hlt">2</span>}(PO{sub 4}){sub 3}, and <span class="hlt">Na</span>{sub <span class="hlt">2</span>}Ni{sub <span class="hlt">2</span>}Cr(PO{sub 4}){sub 3} delivered specific capacities of 352, 385, and 368 mA h g{sup −1}, respectively, which attests to the electrochemical activity of sodium in these compounds. - Highlights: • <span class="hlt">Na</span>CoCr{sub <span class="hlt">2</span>}(PO{sub 4}){sub 3}, <span class="hlt">Na</span>NiCr{sub <span class="hlt">2</span>}(PO{sub 4}){sub 3}, and <span class="hlt">Na</span>{sub <span class="hlt">2</span>}Ni{sub <span class="hlt">2</span>}Cr(PO{sub 4}){sub 3} were synthesized by sol-gel method. • The crystal structures were determined by using neutron powder diffraction data. • The three compounds crystallize with a stuffed α-CrPO{sub 4}-type structure. • The three compounds were tested as anodes in sodium-ion batteries. • Relatively high specific capacities were obtained for these compounds.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://files.eric.ed.gov/fulltext/ED297998.pdf','ERIC'); return false;" href="http://files.eric.ed.gov/fulltext/ED297998.pdf"><span>The Effect of Mode of <span class="hlt">CAI</span> and Individual Learning Differences on the Understanding of Concept Relationships.</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>Rowland, Paul McD.</p> <p></p> <p>The effect of mode of computer-assisted instruction (<span class="hlt">CAI</span>) and individual learning differences on the learning of science concepts was investigated. University elementary education majors learned about home energy use from either a computer simulation or a computer tutorial. Learning of science concepts was measured using achievement and…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20170002376','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20170002376"><span>A New Type of Foreign Clast in A Polymict Ureilite: A <span class="hlt">CAI</span> or AL-Rich Chondrule</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Goodrich, C. A.; Ross, D. K.; Treiman, A. H.</p> <p>2017-01-01</p> <p>Introduction: Polymict ureilites are breccias interpreted to represent regolith formed on a ureilitic asteroid [1-3]. They consist of approximately 90-95% clasts of various ureilite types (olivine-pyroxene rocks with Fo 75-95), a few % indigenous feldspathic clasts, and a few % foreign clasts [4-20]. The foreign clasts are diverse, including fragments of H, L, LL and R chondrites, angrites, other achondrites, and dark clasts similar to CC [6,7,9-19]. We report a new type of foreign clast in polymict ureilite DaG 999. Methods: Clast 8 in Dar al Gani (DaG) 999/1 (Museum fur Naturkunde) was discovered during a survey of feldspathic clasts in polymict ureilites [19,20]. It was studied by BEI, EMPA, and X-ray mapping on the JEOL 8530F electron microprobe at ARES, JSC. Petrography and Mineral Compositions: Clast 8 is sub-rounded to irregular in shape, approximately 85 micrometers in diameter, and consists of approximately 68% pyroxene and 32% mesostasis (by area). Part of the pyroxene (top half of clast in Fig. 1a and <span class="hlt">2</span>) shows a coarse dendritic morphology; the rest appears massive. Mesostasis may be glassy and contains fine needles/grains of pyroxene. The pyroxene has very high CaO (23.5 wt.%) and Al<span class="hlt">2</span>O3 (19.7 wt.%), with the formula: (Ca(0.91)Mg(0.63)Fe(0.01)Al(sup VI) (0.38)Cr(0.01)Ti(0.05)1.99 Si<span class="hlt">2</span>O6. The bulk mesostasis also has very high Al<span class="hlt">2</span>O3 (approximately 26 wt.%). A bulk composition for the clast was obtained by combining modal abundances with phase compositions (Table 1, Fig. 3). Discussion: The pyroxene in clast 8 has a Ca-Al-(Ti)- rich (fassaitic) composition that is clearly distinct from compositions of pyroxenes in main group ureilites [22] or indigenous feldspathic clasts in polymict ureilites [4-8]. It also has significantly higher Al than fassaite in angrites (up to approximately 12 wt.% [23]), which occur as xenoliths in polymict ureilites. Ca-Al-Ti rich pyroxenes are most commonly found in <span class="hlt">CAIs</span>, Al-rich chondrules and other types of refractory</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2005GeCoA..69.5537D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2005GeCoA..69.5537D"><span>Thermodynamic modeling of melts in the system <span class="hlt">Na</span> <span class="hlt">2</span>O-<span class="hlt">Na</span>AlO <span class="hlt">2</span>-SiO <span class="hlt">2</span>-F <span class="hlt">2</span>O -1</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Dolejš, David; Baker, Don R.</p> <p>2005-12-01</p> <p>Fluorine is a common volatile element in magmatic-hydrothermal systems, but its solution mechanisms and thermodynamic description in highly polymerized silicate melts are poorly known. We have developed a thermodynamic model for fluorosilicate liquids that links experimentally determined phase equilibria and spectroscopic information on melt structure. The model is applicable to crystallization of fluoride minerals, fluoride-silicate immiscibility in natural felsic melts, and metallurgical processes. Configurational properties of fluorosilicate melts are described by mixing on three site levels (sublattices): (1) alkali fluoride, polyhedral aluminofluoride and silicofluoride species and nonbridging terminations of the aluminosilicate network, (<span class="hlt">2</span>) alkali-aluminate and silicate tetrahedra within the network and (3) bridging oxygen, nonbridging oxygen and terminal fluorine atoms on tetrahedral apices of the network. Abundances of individual chemical species are described by a homogeneous equilibrium representing melt depolymerization: F - (free) + O 0 (bridging) = F 0 (terminal) + O - (nonbridging) which corresponds to a replacement of an oxygen bridging two tetrahedra by a pair of terminations, one with F and the other with an O and a charge-balancing <span class="hlt">Na</span>. In cryolite-bearing systems two additional interaction mechanisms occur: (1) the self-dissociation of octahedral aluminofluoride complexes: [AlF 6] = [AlF 4] + <span class="hlt">2</span> [F], and (<span class="hlt">2</span>) the short-range order between (O,F)-corners and (Si,<span class="hlt">Na</span>Al)-centers of tetrahedra: Si-O-Si + <span class="hlt">2</span> [<span class="hlt">Na</span>Al]-F = [<span class="hlt">NaAl]-O-[Na</span>Al] + <span class="hlt">2</span> Si-F. Portrayal of these equilibria in ternary Thompson reaction space allows for the decrease in the number of interaction mechanisms by linearly combining melt depolymerization with tetrahedral short-range order. In this formulation, the nonideal thermodynamic properties are represented by reaction energies of homogeneous equilibria, thus defining directly individual chemical species concentrations and configurational</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015JPhD...48G5102L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015JPhD...48G5102L"><span>Experimental and first-principles study of photoluminescent and optical properties of <span class="hlt">Na</span>-doped CuAlO<span class="hlt">2</span>: the role of the <span class="hlt">Na</span>Al-<span class="hlt">2</span><span class="hlt">Na</span> i complex</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Liu, Ruijian; Li, Yongfeng; Yao, Bin; Ding, Zhanhui; Deng, Rui; Zhang, Ligong; Zhao, Haifeng; Liu, Lei</p> <p>2015-08-01</p> <p>We report that a band-tail emission at 3.08 eV, lower than near-band-edge energy, is observed in photoluminescence measurements of bulk <span class="hlt">Na</span>-doped CuAlO<span class="hlt">2</span>. The band-tail emission is attributed to <span class="hlt">Na</span>-related defects. Electronic structure calculations based on the first-principles method demonstrate that the donor-acceptor compensated complex of <span class="hlt">Na</span>Al-<span class="hlt">2</span><span class="hlt">Na</span> i in <span class="hlt">Na</span>-doped CuAlO<span class="hlt">2</span> plays a key role in leading to the band-tail emission and bandgap narrowing. Furthermore, Hall effect measurements indicates that the hole concentration in CuAlO<span class="hlt">2</span> is independent on <span class="hlt">Na</span> doping, which is well understood by the donor-acceptor compensation effect of <span class="hlt">Na</span>Al-<span class="hlt">2</span><span class="hlt">Na</span> i complex.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3366464','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3366464"><span>Intracellular Calcium and the Mechanism of Anodal Supernormal Excitability in Langendorff Perfused Rabbit Ventricles</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Joung, Boyoung; Park, Hyung-Wook; Maruyama, Mitsunori; Tang, Liang; Song, Juan; Han, Seongwook; Piccirillo, Gianfranco; Weiss, James N.; Lin, Shien-Fong; Chen, Peng-Sheng</p> <p>2012-01-01</p> <p>Background Anodal stimulation hyperpolarizes cell membrane and increases intracellular Ca<span class="hlt">2</span>+ (<span class="hlt">Cai</span>) transient. This study tested the hypothesis that The maximum slope of <span class="hlt">Cai</span> decline (–(d<span class="hlt">Cai</span>/dt)max) corresponds to the timing of anodal dip on the strength-interval curve and the initiation of repetitive responses and ventricular fibrillation (VF) after a premature stimulus (S<span class="hlt">2</span>). Methods and Results We simultaneously mapped membrane potential (Vm) and <span class="hlt">Cai</span> in 23 rabbit ventricles. A dip was observed on the anodal strength-interval curve. During the anodal dip, ventricles were captured by anodal break excitation directly under the S<span class="hlt">2</span> electrode. The <span class="hlt">Cai</span> following anodal stimuli is larger than that following cathodal stimuli. The S1-S<span class="hlt">2</span> intervals of the anodal dip (203 ± 10 ms) coincided with the -(d<span class="hlt">Cai</span>/dt)max (199 ± 10 ms, p=NS). BAPTA-AM (n=3), INCX inhibition by low extracellular <span class="hlt">Na</span>+ (n=3), and combined ryanodine and thapsigargin infusion (n=<span class="hlt">2</span>) eliminated the anodal supernormality. Strong S<span class="hlt">2</span> during the relative refractory period (n=5) induced 29 repetitive responses and 10 VF episodes. The interval between S<span class="hlt">2</span> and the first non-driven beat was coincidental with the time of -(d<span class="hlt">Cai</span>/dt)max. Conclusions Larger <span class="hlt">Cai</span> transient and INCX activation induced by anodal stimulation produces anodal supernormality. Time of maximum INCX activation is coincidental to the induction of non- driven beats from the <span class="hlt">Cai</span> sinkhole after a strong premature stimulation. PMID:21301131</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/21301131','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21301131"><span>Intracellular calcium and the mechanism of anodal supernormal excitability in langendorff perfused rabbit ventricles.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Joung, Boyoung; Park, Hyung-Wook; Maruyama, Mitsunori; Tang, Liang; Song, Juan; Han, Seongwook; Piccirillo, Gianfranco; Weiss, James N; Lin, Shien-Fong; Chen, Peng-Sheng</p> <p>2011-01-01</p> <p>Anodal stimulation hyperpolarizes the cell membrane and increases the intracellular Ca(<span class="hlt">2</span>+) (<span class="hlt">Ca(i</span>)) transient. This study tested the hypothesis that the maximum slope of the <span class="hlt">Ca(i</span>) decline (-(dCa(i)/dt)(max)) corresponds to the timing of anodal dip on the strength-interval curve and the initiation of repetitive responses and ventricular fibrillation (VF) after a premature stimulus (S(<span class="hlt">2</span>)). We simultaneously mapped the membrane potential (V(m)) and <span class="hlt">Ca(i</span>) in 23 rabbit ventricles. A dip in the anodal strength-interval curve was observed. During the anodal dip, ventricles were captured by anodal break excitation directly under the S(<span class="hlt">2</span>) electrode. The <span class="hlt">Ca(i</span>) following anodal stimuli is larger than that following cathodal stimuli. The S(1)-S(<span class="hlt">2</span>) intervals of the anodal dip (203±10 ms) coincided with the -(dCa(i)/dt)(max) (199±10 ms, P=NS). BAPTA-AM (n=3), inhibition of the electrogenic <span class="hlt">Na</span>(+)-Ca(<span class="hlt">2</span>+) exchanger current (I(NCX)) by low extracellular <span class="hlt">Na</span>(+) (n=3), and combined ryanodine and thapsigargin infusion (n=<span class="hlt">2</span>) eliminated the anodal supernormality. Strong S(<span class="hlt">2</span>) during the relative refractory period (n=5) induced 29 repetitive responses and 10 VF episodes. The interval between S(<span class="hlt">2</span>) and the first non-driven beat was coincidental with the time of -(dCa(i)/dt)(max). Larger <span class="hlt">Ca(i</span>) transient and I(NCX) activation induced by anodal stimulation produces anodal supernormality. The time of maximum I(NCX) activation is coincidental to the induction of non-driven beats from the <span class="hlt">Ca(i</span>) sinkhole after a strong premature stimulation. All rights are reserved to the Japanese Circulation Society.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28597884','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28597884"><span>A neural network potential energy surface for the <span class="hlt">Na</span>H<span class="hlt">2</span> system and dynamics studies on the H(<span class="hlt">2</span>S) + <span class="hlt">Na</span>H(X1Σ+) → <span class="hlt">Na</span>(<span class="hlt">2</span>S) + H<span class="hlt">2</span>(X1Σg+) reaction.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wang, Shufen; Yuan, Jiuchuang; Li, Huixing; Chen, Maodu</p> <p>2017-08-02</p> <p>In order to study the dynamics of the reaction H( <span class="hlt">2</span> S) + <span class="hlt">Na</span>H(X 1 Σ + ) → <span class="hlt">Na</span>( <span class="hlt">2</span> S) + H <span class="hlt">2</span> (X 1 Σ g + ), a new potential energy surface (PES) for the ground state of the <span class="hlt">Na</span>H <span class="hlt">2</span> system is constructed based on 35 730 ab initio energy points. Using basis sets of quadruple zeta quality, multireference configuration interaction calculations with Davidson correction were carried out to obtain the ab initio energy points. The neural network method is used to fit the PES, and the root mean square error is very small (0.00639 eV). The bond lengths, dissociation energies, zero-point energies and spectroscopic constants of H <span class="hlt">2</span> (X 1 Σ g + ) and <span class="hlt">Na</span>H(X 1 Σ + ) obtained on the new <span class="hlt">Na</span>H <span class="hlt">2</span> PES are in good agreement with the experiment data. On the new PES, the reactant coordinate-based time-dependent wave packet method is applied to study the reaction dynamics of H( <span class="hlt">2</span> S) + <span class="hlt">Na</span>H(X 1 Σ + ) → <span class="hlt">Na</span>( <span class="hlt">2</span> S) + H <span class="hlt">2</span> (X 1 Σ g + ), and the reaction probabilities, integral cross-sections (ICSs) and differential cross-sections (DCSs) are obtained. There is no threshold in the reaction due to the absence of an energy barrier on the minimum energy path. When the collision energy increases, the ICSs decrease from a high value at low collision energy. The DCS results show that the angular distribution of the product molecules tends to the forward direction. Compared with the LiH <span class="hlt">2</span> system, the <span class="hlt">Na</span>H <span class="hlt">2</span> system has a larger mass and the PES has a larger well at the H-<span class="hlt">Na</span>H configuration, which leads to a higher ICS value in the H( <span class="hlt">2</span> S) + <span class="hlt">Na</span>H(X 1 Σ + ) → <span class="hlt">Na</span>( <span class="hlt">2</span> S) + H <span class="hlt">2</span> (X 1 Σ g + ) reaction. Because the H( <span class="hlt">2</span> S) + <span class="hlt">Na</span>H(X 1 Σ + ) → <span class="hlt">Na</span>( <span class="hlt">2</span> S) + H <span class="hlt">2</span> (X 1 Σ g + ) reaction releases more energy, the product molecules can be excited to a higher vibrational state.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013EGUGA..15.4567H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013EGUGA..15.4567H"><span>Europlanet <span class="hlt">NA</span><span class="hlt">2</span> Science Networking</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Harri, Ari-Matti; Szego, Karoly; Genzer, Maria; Schmidt, Walter; Krupp, Norbert; Lammer, Helmut; Kallio, Esa; Haukka, Harri</p> <p>2013-04-01</p> <p>Europlanet RI / <span class="hlt">NA</span><span class="hlt">2</span> Science Networking [1] focused on determining the major goals of current and future European planetary science, relating them to the Research Infrastructure that the Europlanet RI project [<span class="hlt">2</span>] developed, and placing them in a more global context. <span class="hlt">NA</span><span class="hlt">2</span> also enhanced the ability of European planetary scientists to participate on the global scene with their own agenda-setting projects and ideas. The Networking Activity <span class="hlt">NA</span><span class="hlt">2</span> included five working groups, aimed at identifying key science issues and producing reference books on major science themes that will bridge the gap between the results of present and past missions and the scientific preparation of the future ones. Within the Europlanet RI project (2009-2012) the <span class="hlt">NA</span><span class="hlt">2</span> and <span class="hlt">NA</span><span class="hlt">2</span>-WGs organized thematic workshops, an expert exchange program and training groups to improve the scientific impact of this Infrastructure. The principal tasks addressed by <span class="hlt">NA</span><span class="hlt">2</span> were: • Science activities in support to the optimal use of data from past and present space missions, involving the broad planetary science community beyond the "space club" • Science activities in support to the preparation of future planetary missions: Earth-based preparatory observations, laboratory studies, R&D on advanced instrumentation and exploration technologies for the future, theory and modeling etc. • Develop scientific activities, joint publications, dedicated meetings, tools and services, education activities, engaging the public and industries • Update science themes and addressing the two main scientific objectives • Prepare and support workshops of the International Space Science Institute (ISSI) in Bern and • Support Trans National Activities (TNAs), Joined Research Activities (JRAs) and the Integrated and Distributed Information Service (IDIS) of the Europlanet project These tasks were achieved by WG workshops organized by the <span class="hlt">NA</span><span class="hlt">2</span> working groups, by ISSI workshops and by an Expert Exchange Program. There were 17 official WG</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.ncbi.nlm.nih.gov/pubmed/28164158','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28164158"><span>Quasi-solid state rechargeable <span class="hlt">Na</span>-CO<span class="hlt">2</span> batteries with reduced graphene oxide <span class="hlt">Na</span> anodes.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Hu, Xiaofei; Li, Zifan; Zhao, Yaran; Sun, Jianchao; Zhao, Qing; Wang, Jianbin; Tao, Zhanliang; Chen, Jun</p> <p>2017-02-01</p> <p><span class="hlt">Na</span>-CO <span class="hlt">2</span> batteries using earth-abundant <span class="hlt">Na</span> and greenhouse gas CO <span class="hlt">2</span> are promising tools for mobile and stationary energy storage, but they still pose safety risks from leakage of liquid electrolyte and instability of the <span class="hlt">Na</span> metal anode. These issues result in extremely harsh operating conditions of <span class="hlt">Na</span>-CO <span class="hlt">2</span> batteries and increase the difficulty of scaling up this technology. We report the development of quasi-solid state <span class="hlt">Na</span>-CO <span class="hlt">2</span> batteries with high safety using composite polymer electrolyte (CPE) and reduced graphene oxide (rGO) <span class="hlt">Na</span> anodes. The CPE of PVDF-HFP [poly(vinylidene fluoride- co -hexafluoropropylene)]-4% SiO <span class="hlt">2</span> /<span class="hlt">Na</span>ClO 4 -TEGDME (tetraethylene glycol dimethyl ether) has high ion conductivity (1.0 mS cm -1 ), robust toughness, a nonflammable matrix, and strong electrolyte-locking ability. In addition, the rGO-<span class="hlt">Na</span> anode presents fast and nondendritic <span class="hlt">Na</span> + plating/stripping (5.7 to 16.5 mA cm -<span class="hlt">2</span> ). The improved kinetics and safety enable the constructed rGO-<span class="hlt">Na</span>/CPE/CO <span class="hlt">2</span> batteries to successfully cycle in wide CO <span class="hlt">2</span> partial pressure window (5 to 100%, simulated car exhaust) and especially to run for 400 cycles at 500 mA g -1 with a fixed capacity of 1000 mA·hour g -1 in pure CO <span class="hlt">2</span> . Furthermore, we scaled up the reversible capacity to 1.1 A·hour in pouch-type batteries (20 × 20 cm, 10 g, 232 Wh kg -1 ). This study makes quasi-solid state <span class="hlt">Na</span>-CO <span class="hlt">2</span> batteries an attractive prospect.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5287700','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5287700"><span>Quasi–solid state rechargeable <span class="hlt">Na</span>-CO<span class="hlt">2</span> batteries with reduced graphene oxide <span class="hlt">Na</span> anodes</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Hu, Xiaofei; Li, Zifan; Zhao, Yaran; Sun, Jianchao; Zhao, Qing; Wang, Jianbin; Tao, Zhanliang; Chen, Jun</p> <p>2017-01-01</p> <p><span class="hlt">Na</span>-CO<span class="hlt">2</span> batteries using earth-abundant <span class="hlt">Na</span> and greenhouse gas CO<span class="hlt">2</span> are promising tools for mobile and stationary energy storage, but they still pose safety risks from leakage of liquid electrolyte and instability of the <span class="hlt">Na</span> metal anode. These issues result in extremely harsh operating conditions of <span class="hlt">Na</span>-CO<span class="hlt">2</span> batteries and increase the difficulty of scaling up this technology. We report the development of quasi–solid state <span class="hlt">Na</span>-CO<span class="hlt">2</span> batteries with high safety using composite polymer electrolyte (CPE) and reduced graphene oxide (rGO) <span class="hlt">Na</span> anodes. The CPE of PVDF-HFP [poly(vinylidene fluoride-co-hexafluoropropylene)]–4% SiO<span class="hlt">2</span>/<span class="hlt">Na</span>ClO4–TEGDME (tetraethylene glycol dimethyl ether) has high ion conductivity (1.0 mS cm−1), robust toughness, a nonflammable matrix, and strong electrolyte-locking ability. In addition, the rGO-<span class="hlt">Na</span> anode presents fast and nondendritic <span class="hlt">Na</span>+ plating/stripping (5.7 to 16.5 mA cm−<span class="hlt">2</span>). The improved kinetics and safety enable the constructed rGO-<span class="hlt">Na</span>/CPE/CO<span class="hlt">2</span> batteries to successfully cycle in wide CO<span class="hlt">2</span> partial pressure window (5 to 100%, simulated car exhaust) and especially to run for 400 cycles at 500 mA g−1 with a fixed capacity of 1000 mA·hour g−1 in pure CO<span class="hlt">2</span>. Furthermore, we scaled up the reversible capacity to 1.1 A·hour in pouch-type batteries (20 × 20 cm, 10 g, 232 Wh kg−1). This study makes quasi–solid state <span class="hlt">Na</span>-CO<span class="hlt">2</span> batteries an attractive prospect. PMID:28164158</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/20919396-experimental-study-combustion-characteristics-scci-cai-based-direct-injection-gasoline-engine','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/20919396-experimental-study-combustion-characteristics-scci-cai-based-direct-injection-gasoline-engine"><span>An experimental study of the combustion characteristics in SCCI and <span class="hlt">CAI</span> based on direct-injection gasoline engine</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Lee, C.H.; Lee, K.H.</p> <p>2007-08-15</p> <p>Emissions remain a critical issue affecting engine design and operation, while energy conservation is becoming increasingly important. One approach to favorably address these issues is to achieve homogeneous charge combustion and stratified charge combustion at lower peak temperatures with a variable compression ratio, a variable intake temperature and a trapped rate of the EGR using NVO (negative valve overlap). This experiment was attempted to investigate the origins of these lower temperature auto-ignition phenomena with SCCI and <span class="hlt">CAI</span> using gasoline fuel. In case of SCCI, the combustion and emission characteristics of gasoline-fueled stratified-charge compression ignition (SCCI) engine according to intake temperaturemore » and compression ratio was examined. We investigated the effects of air-fuel ratio, residual EGR rate and injection timing on the <span class="hlt">CAI</span> combustion area. In addition, the effect of injection timing on combustion factors such as the start of combustion, its duration and its heat release rate was also investigated. (author)« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29701452','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29701452"><span>Enhancing Catalyzed Decomposition of <span class="hlt">Na</span><span class="hlt">2</span>CO3 with Co<span class="hlt">2</span>MnO x Nanowire-Decorated Carbon Fibers for Advanced <span class="hlt">Na</span>-CO<span class="hlt">2</span> Batteries.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Fang, Cong; Luo, Jianmin; Jin, Chengbin; Yuan, Huadong; Sheng, Ouwei; Huang, Hui; Gan, Yongping; Xia, Yang; Liang, Chu; Zhang, Jun; Zhang, Wenkui; Tao, Xinyong</p> <p>2018-05-23</p> <p>The metal-CO <span class="hlt">2</span> batteries, especially <span class="hlt">Na</span>-CO <span class="hlt">2</span> , batteries come into sight owing to their high energy density, ability for CO <span class="hlt">2</span> capture, and the abundance of sodium resource. Besides the sluggish electrochemical reactions at the gas cathodes and the instability of the electrolyte at a high voltage, the final discharge product <span class="hlt">Na</span> <span class="hlt">2</span> CO 3 is a solid and poor conductor of electricity, which may cause the high overpotential and poor cycle performance for the <span class="hlt">Na</span>-CO <span class="hlt">2</span> batteries. The promotion of decomposition of <span class="hlt">Na</span> <span class="hlt">2</span> CO 3 should be an efficient strategy to enhance the electrochemical performance. Here, we design a facile <span class="hlt">Na</span> <span class="hlt">2</span> CO 3 activation experiment to screen the efficient cathode catalyst for the <span class="hlt">Na</span>-CO <span class="hlt">2</span> batteries. It is found that the Co <span class="hlt">2</span> MnO x nanowire-decorated carbon fibers (CMO@CF) can promote the <span class="hlt">Na</span> <span class="hlt">2</span> CO 3 decomposition at the lowest voltage among all these metal oxide-decorated carbon fiber structures. After assembling the <span class="hlt">Na</span>-CO <span class="hlt">2</span> batteries, the electrodes based on CMO@CF show lower overpotential and better cycling performance compared with the electrodes based on pristine carbon fibers and other metal oxide-modified carbon fibers. We believe this catalyst screening method and the freestanding structure of the CMO@CF electrode may provide an important reference for the development of advanced <span class="hlt">Na</span>-CO <span class="hlt">2</span> batteries.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016GeCoA.183..176H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016GeCoA.183..176H"><span>Microstructural constraints on complex thermal histories of refractory <span class="hlt">CAI</span>-like objects in an amoeboid olivine aggregate from the ALHA77307 CO3.0 chondrite</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Han, Jangmi; Brearley, Adrian J.</p> <p>2016-06-01</p> <p>We have carried out a FIB/TEM study of refractory <span class="hlt">CAI</span>-like objects in one AOA from the ALHA77307 CO3.0 chondrite. The <span class="hlt">CAI</span>-like objects in the AOA consist of a zoned sequence with a spinel-rich core through an intergrowth layer of spinel and Al-Ti-rich diopside to a diopside rim. The spinel-rich core consists of polycrystalline aggregates of spinel and ±minor melilite showing equilibrated grain boundary textures. The intergrowth layer contains fine-grained diopside and spinel with minor anorthite with highly curved and embayed grain boundaries. The diopside rim consists of polycrystalline aggregates of diopside. The compositions of pyroxene change significantly outward from Al-Ti-rich diopside in contact with the spinel-rich core to Al-Ti-poor diopside next to the surrounding olivine of the AOA. Overall microstructural and chemical characteristics suggest that the spinel-rich core formed under equilibrium conditions whereas the intergrowth layer is the result of reactions that occurred under conditions that departed significantly from equilibrium. The remarkable changes in formation conditions of the <span class="hlt">CAI</span>-like objects may have been achieved by transport and injection of refractory objects into a region of a partially-condensed, Ca,Ti-saturated gas which reacted with spinel and melilite to form Al-Ti-rich diopside. Crystallographically-oriented TiO<span class="hlt">2</span> nanoparticles decorate the grain boundaries between spinel grains and between spinel and Al-Ti-rich diopside grains. During the disequilibrium back-reaction of spinel with a partially-condensed, Ca,Ti-saturated gas, metastable TiO<span class="hlt">2</span> nanoparticles may have condensed by an epitaxial nucleation mechanism and grown on the surface of spinel. These TiO<span class="hlt">2</span> nanoparticles are disordered intergrowths of the two TiO<span class="hlt">2</span> polymorphs, anatase and rutile. These nanoparticles are inferred to have nucleated as anatase that underwent partial transformation into rutile. The local presence of the TiO<span class="hlt">2</span> nanoparticles and intergrowth of anatase and</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19870020494','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19870020494"><span><span class="hlt">Na/beta-alumina/Na</span>AlCl4, Cl<span class="hlt">2</span>/C circulating cell</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Cherng, Jing-Yih; Bennion, Douglas N.</p> <p>1987-01-01</p> <p>A study was made of a high specific energy battery based on a sodium negative electrode and a chlorine positive electrode with molten AlCl3-<span class="hlt">Na</span>Cl electrolyte and a solid beta alumina separator. The basic performance of a <span class="hlt">Na</span> beta-alumina <span class="hlt">Na</span>AlCl4, Cl<span class="hlt">2</span>/C circulating cell at 200 C was demonstrated. This cell can be started at 150 C. The use of melting sodium chloroaluminate electrolyte overcomes some of the material problems associated with the high working temperatures of present molten salt systems, such as <span class="hlt">Na</span>/S and LiAl/FeS, and retains the advantages of high energy density and relatively efficient electrode processes. Preliminary investigations were conducted on a sodium-chlorine static cell, material compability, electrode design, wetting, and theoretical calculations to assure a better chance of success before assembling a <span class="hlt">Na</span>/Cl<span class="hlt">2</span> circulating cell. Mathematical models provide a theoretical explanation for the performance of the <span class="hlt">Na</span>Cl<span class="hlt">2</span> battery. The results of mathematical models match the experimental results very well. According to the result of the mathematical modeling, an output at 180 mA/sq cm and 3.<span class="hlt">2</span> V can be obtained with optimized cell design.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://files.eric.ed.gov/fulltext/ED077195.pdf','ERIC'); return false;" href="http://files.eric.ed.gov/fulltext/ED077195.pdf"><span>A Model Driven Question-Answering System for a <span class="hlt">CAI</span> Environment. Final Report (July 1970 to May 1972).</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>Brown, John S.; And Others</p> <p></p> <p>A question answering system which permits a computer-assisted instruction (<span class="hlt">CAI</span>) student greater initiative in the variety of questions he can ask is described. A method is presented to represent the dynamic processes of a subject matter area by augmented finite state automata, which permits efficient inferencing about dynamic processes and…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1998JCP....95.1711M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1998JCP....95.1711M"><span>Dynamic polarizabilities and Van der Waals coefficients for alkali atoms Li, <span class="hlt">Na</span> and alkali dimer molecules Li<span class="hlt">2</span>, <span class="hlt">Na</span><span class="hlt">2</span> and <span class="hlt">Na</span>Li</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mérawa, M.; Dargelos, A.</p> <p>1998-07-01</p> <p>The present paper gives an account of investigations of the polarizability of the alkali atoms Li, <span class="hlt">Na</span>, diatomics homonuclear and heteronuclear Li<span class="hlt">2</span>, <span class="hlt">Na</span><span class="hlt">2</span> and <span class="hlt">Na</span>Li at SCF (Self Consistent Field) level of approximation and at correlated level, using a time Time-Dependent Gauge Invariant method (TDGI). Our static polarizability values agree with the best experimental and theoretical determinations. The Van der Waals C6 coefficients for the atom-atom, atom-dimer and dimer-dimer interactions have been evaluated. Les polarisabilités des atomes alcalins Li, <span class="hlt">Na</span>, et des molécules diatomiques homonucléaires et hétéronucléaire Li<span class="hlt">2</span>, <span class="hlt">Na</span><span class="hlt">2</span> et <span class="hlt">Na</span>Li, ont été calculées au niveau SCF (Self Consistent Field) et au niveau corrélé à partir d'une méthode invariante de jauge dépendante du temps(TDGI). Nos valeurs des polarisabilités statiques sont en accord avec les meilleurs déterminations expérimentales et théoriques. Les coefficients C6 de Van de Waals pour les interactions atome-atome, atome-dimère et dimère-dimère ont également été évalués.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19880051937&hterms=CO2+H2O&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3DCO2%2BH2O','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19880051937&hterms=CO2+H2O&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3DCO2%2BH2O"><span>Infrared reflectance spectra of <span class="hlt">Na</span><span class="hlt">2</span>S with contaminant <span class="hlt">Na</span><span class="hlt">2</span>CO3 - Effects of adsorbed H<span class="hlt">2</span>O and CO<span class="hlt">2</span> and relation to studies of Io</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Nash, Douglas B.</p> <p>1988-01-01</p> <p>A previously reported laboratory determination of the IR spectrum of <span class="hlt">Na</span><span class="hlt">2</span>S is presently noted to have been incorrectly interpreted, due to the inadvertent contamination of the sample with <span class="hlt">Na</span><span class="hlt">2</span>CO3. New <span class="hlt">Na</span><span class="hlt">2</span>S spectra are presented, and the <span class="hlt">Na</span><span class="hlt">2</span>CO3 spectrum is examined in order to demonstrate that this phase is the primary sample contaminant. <span class="hlt">Na</span><span class="hlt">2</span>S is a candidate surface component on the Jupiter satellite, Io, in view of its apparent high IR brightness and spectral neutrality in the 1-5 micron range.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27693267','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27693267"><span>Complete genome sequence of Defluviimonas alba <span class="hlt">cai</span>42T, a microbial exopolysaccharides producer.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zhao, Jie-Yu; Geng, Shuang; Xu, Lian; Hu, Bing; Sun, Ji-Quan; Nie, Yong; Tang, Yue-Qin; Wu, Xiao-Lei</p> <p>2016-12-10</p> <p>Defluviimonas alba <span class="hlt">cai</span>42 T , isolated from the oil-production water in Xinjiang Oilfield in China, has a strong ability to produce exopolysaccharides (EPS). We hereby present its complete genome sequence information which consists of a circular chromosome and three plasmids. The strain characteristically contains various genes encoding for enzymes involved in EPS biosynthesis, modification, and export. According to the genomic and physiochemical data, it is predicted that the strain has the potential to be utilized in industrial production of microbial EPS. Copyright © 2016 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19830039833&hterms=K2&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3DK2','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19830039833&hterms=K2&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3DK2"><span>A semiclassical study of laser-induced atomic fluorescence from <span class="hlt">Na</span><span class="hlt">2</span>, K<span class="hlt">2</span> and <span class="hlt">Na</span>K</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Yuan, J.-M.; Bhattacharyya, D. K.; George, T. F.</p> <p>1982-01-01</p> <p>A semiclassical treatment of laser-induced atomic fluorescence for the alkali-dimer systems <span class="hlt">Na</span><span class="hlt">2</span>, K<span class="hlt">2</span> and <span class="hlt">Na</span>K is presented. The variation of the fluorescence intensity with the frequency of the exciting laser photon is studied and a comparison of theoretical results with a set of experimental data is presented.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://eric.ed.gov/?q=poster+AND+presentation+AND+psychology&pg=4&id=ED383534','ERIC'); return false;" href="https://eric.ed.gov/?q=poster+AND+presentation+AND+psychology&pg=4&id=ED383534"><span>Proceedings of the Annual Meeting of the North American Chapter of the International Group for the Psychology of Mathematics Education (16th, Baton Rouge, Louisiana, November 5-8, 1994). Volume <span class="hlt">2</span>: Research Papers, Oral Reports, and Posters (Continued).</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>Kirshner, David, Ed.</p> <p></p> <p>This PME-<span class="hlt">NA</span> proceedings volume contains the full text of 41 research papers. In addition, brief usually one-page reports, are provided for 11 oral presentations and 13 poster sessions. The full research reports are as follows: "Cognitive Analysis of Chinese Students' Mathematical Problem Solving" (J. <span class="hlt">Cai</span> and E. A. Silver); (<span class="hlt">2</span>)…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1437023-relative-viscosity-nano3-nano2-aqueous-solutions','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1437023-relative-viscosity-nano3-nano2-aqueous-solutions"><span>The relative viscosity of <span class="hlt">Na</span>NO 3 and <span class="hlt">Na</span>NO <span class="hlt">2</span> aqueous solutions</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Reynolds, Jacob G.; Mauss, Billie M.; Daniel, Richard C.</p> <p>2018-05-09</p> <p>In aqueous solution, both nitrate and nitrite are planar, monovalent, and have the same elements but different sizes and charge densities. Comparing the viscosity of <span class="hlt">Na</span>NO <span class="hlt">2</span> and <span class="hlt">Na</span>NO 3 aqueous solutions provides an opportunity to determine the relative importance of anion size versus strength of anion interaction with water. The viscosity of aqueous <span class="hlt">Na</span>NO <span class="hlt">2</span> and <span class="hlt">Na</span>NO 3 were measured over a temperature and concentration range relevant to nuclear waste processing. The viscosity of <span class="hlt">Na</span>NO <span class="hlt">2</span> solutions was consistently larger than <span class="hlt">Na</span>NO 3 under all conditions, even though nitrate is larger than nitrite. This was interpreted in terms ofmore » quantum mechanical charge field molecular dynamics calculations that indicate that nitrite forms more and stronger hydrogen bonds with water per oxygen atom than nitrate. Furthermore, these hydrogen bonds inhibit rotational motion required for fluid flow, thus increasing the nitrite solution viscosity relative to that of an equivalent nitrate solution.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/1437023-relative-viscosity-nano3-nano2-aqueous-solutions','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/1437023-relative-viscosity-nano3-nano2-aqueous-solutions"><span>The relative viscosity of <span class="hlt">Na</span>NO 3 and <span class="hlt">Na</span>NO <span class="hlt">2</span> aqueous solutions</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>Reynolds, Jacob G.; Mauss, Billie M.; Daniel, Richard C.</p> <p></p> <p>In aqueous solution, both nitrate and nitrite are planar, monovalent, and have the same elements but different sizes and charge densities. Comparing the viscosity of <span class="hlt">Na</span>NO <span class="hlt">2</span> and <span class="hlt">Na</span>NO 3 aqueous solutions provides an opportunity to determine the relative importance of anion size versus strength of anion interaction with water. The viscosity of aqueous <span class="hlt">Na</span>NO <span class="hlt">2</span> and <span class="hlt">Na</span>NO 3 were measured over a temperature and concentration range relevant to nuclear waste processing. The viscosity of <span class="hlt">Na</span>NO <span class="hlt">2</span> solutions was consistently larger than <span class="hlt">Na</span>NO 3 under all conditions, even though nitrate is larger than nitrite. This was interpreted in terms ofmore » quantum mechanical charge field molecular dynamics calculations that indicate that nitrite forms more and stronger hydrogen bonds with water per oxygen atom than nitrate. Furthermore, these hydrogen bonds inhibit rotational motion required for fluid flow, thus increasing the nitrite solution viscosity relative to that of an equivalent nitrate solution.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3372088','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3372088"><span>Acidosis Differentially Modulates Inactivation in <span class="hlt">Na</span>V1.<span class="hlt">2</span>, <span class="hlt">Na</span>V1.4, and <span class="hlt">Na</span>V1.5 Channels</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Vilin, Yury Y.; Peters, Colin H.; Ruben, Peter C.</p> <p>2012-01-01</p> <p><span class="hlt">Na</span>V channels play a crucial role in neuronal and muscle excitability. Using whole-cell recordings we studied effects of low extracellular pH on the biophysical properties of <span class="hlt">Na</span>V1.<span class="hlt">2</span>, <span class="hlt">Na</span>V1.4, and <span class="hlt">Na</span>V1.5, expressed in cultured mammalian cells. Low pH produced different effects on different channel subtypes. Whereas <span class="hlt">Na</span>V1.4 exhibited very low sensitivity to acidosis, primarily limited to partial block of macroscopic currents, the effects of low pH on gating in <span class="hlt">Na</span>V1.<span class="hlt">2</span> and <span class="hlt">Na</span>V1.5 were profound. In <span class="hlt">Na</span>V1.<span class="hlt">2</span> low pH reduced apparent valence of steady-state fast inactivation, shifted the τ(V) to depolarizing potentials and decreased channels availability during onset to slow and use-dependent inactivation (UDI). In contrast, low pH delayed open-state inactivation in <span class="hlt">Na</span>V1.5, right-shifted the voltage-dependence of window current, and increased channel availability during onset to slow and UDI. These results suggest that protons affect channel availability in an isoform-specific manner. A computer model incorporating these results demonstrates their effects on membrane excitability. PMID:22701426</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013JCrGr.377...66Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013JCrGr.377...66Y"><span>Formation of Si grains from a <span class="hlt">Na</span>Si melt prepared by reaction of SiO<span class="hlt">2</span> and <span class="hlt">Na</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yamane, Hisanori; Morito, Haruhiko; Uchikoshi, Masahito</p> <p>2013-08-01</p> <p>A mixture of <span class="hlt">Na</span><span class="hlt">2</span>SiO3 and <span class="hlt">Na</span>Si was found to be formed by reaction of SiO<span class="hlt">2</span> and <span class="hlt">Na</span> at 650 °C as follows: 5<span class="hlt">Na</span>+3SiO<span class="hlt">2</span>→<span class="hlt">2</span><span class="hlt">Na</span><span class="hlt">2</span>SiO3+<span class="hlt">Na</span>Si. Single crystals of <span class="hlt">Na</span>Si were grown by cooling the mixture of <span class="hlt">Na</span><span class="hlt">2</span>SiO3 and <span class="hlt">Na</span>Si with an excess amount of <span class="hlt">Na</span> from 850 °C, and polycrystalline Si was obtained by vaporization of <span class="hlt">Na</span> from the crystals. Coarse grains of Si were also crystallized by <span class="hlt">Na</span> evaporation after the formation of <span class="hlt">Na</span><span class="hlt">2</span>SiO3 and Si-dissolved liquid <span class="hlt">Na</span> at 830 °C. The Si grains were collected by washing the product with water. The yield of the Si grains was 85% of the ideal amount expected from the reaction.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.dtic.mil/docs/citations/ADA239997','DTIC-ST'); return false;" href="http://www.dtic.mil/docs/citations/ADA239997"><span>A Design of Computer Aided Instructions (<span class="hlt">CAI</span>) for Undirected Graphs in the Discrete Math Tutorial (DMT). Part 1.</span></a></p> <p><a target="_blank" href="http://www.dtic.mil/">DTIC Science & Technology</a></p> <p></p> <p>1990-06-01</p> <p>The objective of this thesis research is to create a tutorial for teaching aspects of undirected graphs in discrete math . It is one of the submodules...of the Discrete Math Tutorial (DMT), which is a Computer Aided Instructional (<span class="hlt">CAI</span>) tool for teaching discrete math to the Naval Academy and the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24165016','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24165016"><span>Sexual life and sexual wellness in individuals with complete androgen insensitivity syndrome (<span class="hlt">CAIS</span>) and Mayer-Rokitansky-Küster-Hauser Syndrome (MRKHS).</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Fliegner, Maike; Krupp, Kerstin; Brunner, Franziska; Rall, Katharina; Brucker, Sara Y; Briken, Peer; Richter-Appelt, Hertha</p> <p>2014-03-01</p> <p>Sexual wellness depends on a person's physical and psychological constitution. Complete Androgen Insensitivity Syndrome (<span class="hlt">CAIS</span>) and Mayer-Rokitansky-Küster-Hauser Syndrome (MRKHS) can compromise sexual well-being. To compare sexual well-being in <span class="hlt">CAIS</span> and MRKHS using multiple measures: To assess sexual problems and perceived distress. To gain insight into participants' feelings of inadequacy in social and sexual situations, level of self-esteem and depression. To determine how these psychological factors relate to sexual (dys)function. To uncover what participants see as the source of their sexual problems. Data were collected using a paper-and-pencil questionnaire. Eleven individuals with <span class="hlt">CAIS</span> and 49 with MRKHS with/without neovagina treatment were included. Rates of sexual dysfunctions, overall sexual function, feelings of inadequacy in social and sexual situations, self-esteem and depression scores were calculated. Categorizations were used to identify critical cases. Correlations between psychological variables and sexual function were computed. Sexually active subjects were compared with sexually not active participants. A qualitative content analysis was carried out to explore causes of sexual problems. An extended list of sexual problems based on the Diagnostic and Statistical Manual of Mental Disorders, 4th ed., text revision, by the American Psychiatric Association and related distress. Female Sexual Function Index (FSFI), German Questionnaire on Feelings of Inadequacy in Social and Sexual Situations (FUSS social scale, FUSS sexual scale), Rosenberg Self-Esteem Scale (RSE), Brief Symptom Inventory (BSI) subscale depression. Open question on alleged causes of sexual problems. The results point to a far-reaching lack of sexual confidence and sexual satisfaction in <span class="hlt">CAIS</span>. In MRKHS apprehension in sexual situations is a source of distress, but sexual problems seem to be more focused on issues of vaginal functioning. MRKHS women report being satisfied with their</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016CP....475..131T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016CP....475..131T"><span>Crystallization kinetics from mixture <span class="hlt">Na</span><span class="hlt">2</span>SO4/glycerol droplets of <span class="hlt">Na</span><span class="hlt">2</span>SO4 by FTIR-ATR</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Tan, Dan-Ting; Cai, Chen; Zhang, Yun; Wang, Na; Pang, Shu-Feng; Zhang, Yun-Hong</p> <p>2016-08-01</p> <p>The efflorescence of mixed <span class="hlt">Na</span><span class="hlt">2</span>SO4/glycerol aerosols on the ZnSe substrate with various mole ratios (<span class="hlt">Na</span><span class="hlt">2</span>SO4/glycerol = 1:1, 1:<span class="hlt">2</span>, 1:4) has been studied in the relative humidity (RH) linearly decline process, using a situ Fourier transform infrared attenuated total reflection (FTIR-ATR) technique. The crystal ratio at a given RH can be gained by the absorbance of the band at 1132 cm-1, which shows the incomplete nucleation for mixed <span class="hlt">Na</span><span class="hlt">2</span>SO4/glycerol aerosols and the decreased amount of the droplets crystallized at the lowest RH with the glycerol increase. Using the volume fraction of droplets that have yet to crystallize, the heterogeneous nucleation kinetics has been gained. By the Extended Aerosol Inorganics Model (E-AIM), the nucleation rate as the function of solute saturation degree has been gained for various mixed <span class="hlt">Na</span><span class="hlt">2</span>SO4/glycerol aerosols.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27666995','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27666995"><span>Effects of Lactobacillus curvatus and Leuconostoc mesenteroides on Suan <span class="hlt">Cai</span> Fermentation in Northeast China.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Yang, Hongyan; Wu, Hao; Gao, Lijuan; Jia, Hongbai; Zhang, Yuan; Cui, Zongjun; Li, Yuhua</p> <p>2016-12-28</p> <p>To investigate the effects of Lactobacillus curvatus and Leuconostoc mesenteroides on suan <span class="hlt">cai</span> (pickled Chinese cabbage) fermentation, L. curvatus and/or Ln. mesenteroides were inoculated into suan <span class="hlt">cai</span>. Physicochemical indexes were measured, and the microbial dynamics during the fermentation were analyzed by Illumina MiSeq sequencing and quantitative polymerase chain reaction (qPCR). The results showed that inoculation with lactic acid bacteria (LAB) lowered the pH of the fermentation system more rapidly. The decrease in water-soluble carbohydrates in the inoculated treatments occurred more rapidly than in the control. The LAB counts in the control were lower than in other inoculated treatments during the first 12 days of fermentation. According to the Illumina MiSeq sequencing analyses, Firmicutes , Proteobacteria , Bacteroidetes , Actinobacteria , Cyanobacteria , Fusobacteria , and Verrucomicrobia were present in the fermentations, along with other unclassified bacteria. Generally, Firmicutes was predominant during the fermentation in all treatments. At the genus level, 16 genera were detected. The relative abundance of Lactobacillus in all inoculated treatments was higher than in the control. The relative abundance of Lactobacillus in the treatments containing L. curvatus was higher than in the Ln. mesenteroides -only treatment. The relative abundance of Leuconostoc in the Ln. mesenteroides -containing treatments increased continuously throughout the fermentation. Leuconostoc was highest in the Ln. mesenteroides -only treatment. According to the qPCR results, L. curvatus and/or Ln. mesenteroides inoculations could effectively inhabit the fermentation system. L. curvatus dominated the fermentation in the inoculated treatments.</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/26967192','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26967192"><span>Co-intercalation of Mg(<span class="hlt">2</span>+) and <span class="hlt">Na</span>(+) in <span class="hlt">Na</span>(0.69)Fe<span class="hlt">2</span>(CN)6 as a High-Voltage Cathode for Magnesium Batteries.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kim, Dong-Min; Kim, Youngjin; Arumugam, Durairaj; Woo, Sang Won; Jo, Yong Nam; Park, Min-Sik; Kim, Young-Jun; Choi, Nam-Soon; Lee, Kyu Tae</p> <p>2016-04-06</p> <p>Thanks to the advantages of low cost and good safety, magnesium metal batteries get the limelight as substituent for lithium ion batteries. However, the energy density of state-of-the-art magnesium batteries is not high enough because of their low operating potential; thus, it is necessary to improve the energy density by developing new high-voltage cathode materials. In this study, nanosized Berlin green Fe<span class="hlt">2</span>(CN)6 and Prussian blue <span class="hlt">Na</span>(0.69)Fe<span class="hlt">2</span>(CN)6 are compared as high-voltage cathode materials for magnesium batteries. Interestingly, while Mg(<span class="hlt">2</span>+) ions cannot be intercalated in Fe<span class="hlt">2</span>(CN)6, <span class="hlt">Na</span>(0.69)Fe<span class="hlt">2</span>(CN)6 shows reversible intercalation and deintercalation of Mg(<span class="hlt">2</span>+) ions, although they have the same crystal structure except for the presence of <span class="hlt">Na</span>(+) ions. This phenomenon is attributed to the fact that Mg(<span class="hlt">2</span>+) ions are more stable in <span class="hlt">Na</span>(+)-containing <span class="hlt">Na</span>(0.69)Fe<span class="hlt">2</span>(CN)6 than in <span class="hlt">Na</span>(+)-free Fe<span class="hlt">2</span>(CN)6, indicating <span class="hlt">Na</span>(+) ions in <span class="hlt">Na</span>(0.69)Fe<span class="hlt">2</span>(CN)6 plays a crucial role in stabilizing Mg(<span class="hlt">2</span>+) ions. <span class="hlt">Na</span>(0.69)Fe<span class="hlt">2</span>(CN)6 delivers reversible capacity of approximately 70 mA h g(-1) at 3.0 V vs Mg/Mg(<span class="hlt">2</span>+) and shows stable cycle performance over 35 cycles. Therefore, Prussian blue analogues are promising structures for high-voltage cathode materials in Mg batteries. Furthermore, this co-intercalation effect suggests new avenues for the development of cathode materials in hybrid magnesium batteries that use both Mg(<span class="hlt">2</span>+) and <span class="hlt">Na</span>(+) ions as charge carriers.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19890054072&hterms=Thermoelectric+effect&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3DThermoelectric%2Beffect','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19890054072&hterms=Thermoelectric+effect&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3DThermoelectric%2Beffect"><span>Effects of <span class="hlt">Na</span><span class="hlt">2</span>MoO4 and <span class="hlt">Na</span><span class="hlt">2</span>WO4 on molybdenum and tungsten electrodes for the alkali metal thermoelectric converter (AMTEC)</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Williams, R. M.; Wheeler, B. L.; Jeffries-Nakamura, B.; Loveland, M. E.; Bankston, C. P.</p> <p>1988-01-01</p> <p>The effects of adding <span class="hlt">Na</span><span class="hlt">2</span>MoO4 and <span class="hlt">Na</span><span class="hlt">2</span>WO4 to porous Mo and W electrodes, respectively, on the performance and impedance characteristics of the electrodes in an alkali metal thermoelectric converter (AMTEC) were investigated. It was found that corrosion of the porous electrode by <span class="hlt">Na</span><span class="hlt">2</span>MoO4 or <span class="hlt">Na</span><span class="hlt">2</span>WO4 to form <span class="hlt">Na</span><span class="hlt">2</span>MO3O6 and WO<span class="hlt">2</span>, respectively, and recrystallization of the Mo or W as the salt evaporates, result in major morphological changes including a loss of columnar structure and a significant increase in porosity. This effect is more pronounced in <span class="hlt">Na</span><span class="hlt">2</span>MoO4/Mo electrodes, due to the lower stability of <span class="hlt">Na</span><span class="hlt">2</span>MoO4.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://eric.ed.gov/?q=csr+OR+corporate+AND+social+AND+responsibility&id=EJ1062828','ERIC'); return false;" href="https://eric.ed.gov/?q=csr+OR+corporate+AND+social+AND+responsibility&id=EJ1062828"><span>From Corporate Social Responsibility, through Entrepreneurial Orientation, to Knowledge Sharing: A Study in <span class="hlt">Cai</span> Luong (Renovated Theatre) Theatre Companies</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>Tuan, Luu Trong</p> <p>2015-01-01</p> <p>Purpose: This paper aims to examine the role of antecedents such as corporate social responsibility (CSR) and entrepreneurial orientation in the chain effect to knowledge sharing among members of <span class="hlt">Cai</span> Luong theatre companies in the Vietnamese context. Knowledge sharing contributes to the depth of the knowledge pool of both the individuals and the…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013MMI....19.1283L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013MMI....19.1283L"><span>Role of alkali carbonate and salt in topochemical synthesis of K1/<span class="hlt">2</span><span class="hlt">Na</span>1/<span class="hlt">2</span>NbO3 and <span class="hlt">Na</span>NbO3 templates</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lee, Jae-Seok; Jeon, Jae-Ho; Choi, Si-Young</p> <p>2013-11-01</p> <p>Since the properties of lead-free piezoelectric materials have thus far failed to meet those of lead-based materials, either chemical doping or morphological texturing should be employed to improve the piezoelectric properties of lead-free piezoelectric ceramics. The goal of this study was to synthesize plate-like K1/<span class="hlt">2</span><span class="hlt">Na</span>1/<span class="hlt">2</span>NbO3 and <span class="hlt">Na</span>NbO3 particles, which are the most favorable templates for morphological texturing of K1/<span class="hlt">2</span><span class="hlt">Na</span>1/<span class="hlt">2</span>NbO3 ceramics. To achieve this goal, Bi<span class="hlt">2</span>.5<span class="hlt">Na</span>3.5Nb5O18 precursors in a plate-like shape were first synthesized and subsequently converted into K1/<span class="hlt">2</span><span class="hlt">Na</span>1/<span class="hlt">2</span>NbO3 or <span class="hlt">Na</span>NbO3 particles that retain the morphology of Bi<span class="hlt">2</span>.5<span class="hlt">Na</span>3.5Nb5O18. In this study, we found that sodium or potassium carbonate does not play a major role in converting the Bi<span class="hlt">2</span>.5<span class="hlt">Na</span>3.5Nb5O18 precursor to K1/<span class="hlt">2</span><span class="hlt">Na</span>1/<span class="hlt">2</span>NbO3 or <span class="hlt">Na</span>NbO3, on the contrary to previous reports; however, the salt contributes to the conversion reaction. All synthesis processes have been performed via a molten salt method, and scanning electron microscopy, scanning probe microscopy, and inductively coupled plasma mass spectroscopy were used to characterize the synthesized K1/<span class="hlt">2</span><span class="hlt">Na</span>1/<span class="hlt">2</span>NbO3 or <span class="hlt">Na</span>NbO3 templates.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JPS...353...85B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JPS...353...85B"><span>High performance sodium-ion hybrid capacitor based on <span class="hlt">Na</span><span class="hlt">2</span>Ti<span class="hlt">2</span>O4(OH)<span class="hlt">2</span> nanostructures</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Babu, Binson; Shaijumon, M. M.</p> <p>2017-06-01</p> <p>Hybrid <span class="hlt">Na</span>-ion capacitors bridge the performance gap between <span class="hlt">Na</span>-ion batteries and supercapacitors and offer excellent energy and power characteristics. However, designing efficient anode and cathode materials with improved kinetics and long cycle life is essential for practical implementation of this technology. Herein, layered sodium titanium oxide hydroxide, <span class="hlt">Na</span><span class="hlt">2</span>Ti<span class="hlt">2</span>O4(OH)<span class="hlt">2</span>, synthesized through hydrothermal technique, is studied as efficient anode material for hybrid <span class="hlt">Na</span>-ion capacitor. Half-cell electrochemical studies vs. <span class="hlt">Na/Na</span>+ showed excellent performance for <span class="hlt">Na</span><span class="hlt">2</span>Ti<span class="hlt">2</span>O4(OH)<span class="hlt">2</span> electrode, with ∼57.<span class="hlt">2</span>% of the total capacity (323.3 C g-1 at 1.0 mV s-1) dominated by capacitive behavior and the remaining due to <span class="hlt">Na</span>-intercalation. The obtained values are in good agreement with Trasatti plots indicating the potential of this material as efficient anode for hybrid <span class="hlt">Na</span>-ion capacitor. Further, a full cell <span class="hlt">Na</span>-ion capacitor is fabricated with <span class="hlt">Na</span><span class="hlt">2</span>Ti<span class="hlt">2</span>O4(OH)<span class="hlt">2</span> as anode and chemically activated Rice Husk Derived Porous Carbon (RHDPC-KOH) as cathode by using organic electrolyte. The hybrid device, operated at a maximum cell voltage of 4 V, exhibits stable electrochemical performance with a maximum energy density of ∼65 Wh kg-1 (at 500 W kg-1, 0.20 A g-1) and with more than ∼ 93% capacitive retention after 3000 cycles.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2009gdca.conf..321C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009gdca.conf..321C"><span>Problem Solving Process Research of Everyone Involved in Innovation Based on <span class="hlt">CAI</span> Technology</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Chen, Tao; Shao, Yunfei; Tang, Xiaowo</p> <p></p> <p>It is very important that non-technical department personnel especially bottom line employee serve as innovators under the requirements of everyone involved in innovation. According the view of this paper, it is feasible and necessary to build everyone involved in innovation problem solving process under Total Innovation Management (TIM) based on the Theory of Inventive Problem Solving (TRIZ). The tools under the <span class="hlt">CAI</span> technology: How TO mode and science effects database could be very useful for all employee especially non-technical department and bottom line for innovation. The problem solving process put forward in the paper focus on non-technical department personnel especially bottom line employee for innovation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28462412','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28462412"><span>Mechanistic origin of low polarization in aprotic <span class="hlt">Na</span>-O<span class="hlt">2</span> batteries.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Ma, Shunchao; McKee, William C; Wang, Jiawei; Guo, Limin; Jansen, Martin; Xu, Ye; Peng, Zhangquan</p> <p>2017-05-21</p> <p>Research interest in aprotic sodium-air (<span class="hlt">Na</span>-O <span class="hlt">2</span> ) batteries is growing because of their considerably high theoretical specific energy and potentially better reversibility than lithium-air (Li-O <span class="hlt">2</span> ) batteries. While Li <span class="hlt">2</span> O <span class="hlt">2</span> has been unequivocally identified as the major discharge product in Li-O <span class="hlt">2</span> batteries containing relatively stable electrolytes, a multitude of discharge products, including <span class="hlt">Na</span>O <span class="hlt">2</span> , <span class="hlt">Na</span> <span class="hlt">2</span> O <span class="hlt">2</span> and <span class="hlt">Na</span> <span class="hlt">2</span> O <span class="hlt">2</span> ·<span class="hlt">2</span>H <span class="hlt">2</span> O, have been reported for <span class="hlt">Na</span>-O <span class="hlt">2</span> batteries and the corresponding cathodic electrochemistry remains incompletely understood. Herein, we provide molecular-level insights into the key mechanistic differences between <span class="hlt">Na</span>-O <span class="hlt">2</span> and Li-O <span class="hlt">2</span> batteries based on gold electrodes in strictly dry, aprotic dimethyl sulfoxide electrolytes through a combination of in situ spectroelectrochemistry and density functional theory based modeling. While like Li-O <span class="hlt">2</span> batteries, the formation of oxygen reduction products (i.e., O <span class="hlt">2</span> - , <span class="hlt">Na</span>O <span class="hlt">2</span> and <span class="hlt">Na</span> <span class="hlt">2</span> O <span class="hlt">2</span> ) in <span class="hlt">Na</span>-O <span class="hlt">2</span> batteries depends critically on the electrode potential, two factors lead to a better reversibility of <span class="hlt">Na</span>-O <span class="hlt">2</span> electrochemistry, and are therefore highly beneficial to a viable rechargeable metal-air battery design: (i) only O <span class="hlt">2</span> - and <span class="hlt">Na</span>O <span class="hlt">2</span> , and no <span class="hlt">Na</span> <span class="hlt">2</span> O <span class="hlt">2</span> , form down to as low as ∼1.5 V vs. <span class="hlt">Na/Na</span> + during discharge; (ii) solid <span class="hlt">Na</span>O <span class="hlt">2</span> is quite soluble and its formation and oxidation can proceed through micro-reversible EC (a chemical reaction of the product after the electron transfer) and CE (a chemical reaction preceding the electron transfer) processes, respectively, with O <span class="hlt">2</span> - as the key intermediate.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://files.eric.ed.gov/fulltext/ED069154.pdf','ERIC'); return false;" href="http://files.eric.ed.gov/fulltext/ED069154.pdf"><span>The Computer as an Authority Figure: Some Effects of <span class="hlt">CAI</span> on Student Perception of Teacher Authority. Technical Report Number 29.</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>Brod, Rodney L.</p> <p></p> <p>A sociological theory of authority was used to investigate some nonintellective, perhaps unintended, consequences of computer-assisted instruction (<span class="hlt">CAI</span>) upon student's attitudes and orientations toward the organization of the school. An attitudinal questionnaire was used to survey attitudes toward the teacher and the computer in a junior high…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018EPJWC.17901009L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018EPJWC.17901009L"><span><span class="hlt">NA</span>62 and <span class="hlt">NA</span>48/<span class="hlt">2</span> results on search for Heavy Neutral Leptons</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lamanna, Gianluca; Aliberti, R.; Ambrosino, F.; Ammendola, R.; Angelucci, B.; Antonelli, A.; Anzivino, G.; Arcidiacono, R.; Barbanera, M.; Biagioni, A.; Bician, L.; Biino, C.; Bizzeti, A.; Blazek, T.; Bloch-Devaux, B.; Bonaiuto, V.; Boretto, M.; Bragadireanu, M.; Britton, D.; Brizioli, F.; Brunetti, M. B.; Bryman, D.; Bucci, F.; Capussela, T.; Ceccucci, A.; Cenci, P.; Cerny, V.; Cerri, C.; Checcucci, B.; Conovaloff, A.; Cooper, P.; Cortina Gil, E.; Corvino, M.; Costantini, F.; Cotta Ramusino, A.; Coward, D.; D'Agostini, G.; Dainton, J.; Dalpiaz, P.; Danielsson, H.; De Simone, N.; Di Filippo, D.; Di Lella, L.; Doble, N.; Dobrich, B.; Duval, F.; Duk, V.; Engelfried, J.; Enik, T.; Estrada-Tristan, N.; Falaleev, V.; Fantechi, R.; Fascianelli, V.; Federici, L.; Fedotov, S.; Filippi, A.; Fiorini, M.; Fry, J.; Fu, J.; Fucci, A.; Fulton, L.; Gamberini, E.; Gatignon, L.; Georgiev, G.; Ghinescu, S.; Gianoli, A.; Giorgi, M.; Giudici, S.; Gonnella, F.; Goudzovski, E.; Graham, C.; Guida, R.; Gushchin, E.; Hahn, F.; Heath, H.; Husek, T.; Hutanu, O.; Hutchcroft, D.; Iacobuzio, L.; Iacopini, E.; Imbergamo, E.; Jenninger, B.; Kampf, K.; Kekelidze, V.; Kholodenko, S.; Khoriauli, G.; Khotyantsev, A.; Kleimenova, A.; Korotkova, A.; Koval, M.; Kozhuharov, V.; Kucerova, Z.; Kudenko, Y.; Kunze, J.; Kurochka, V.; Kurshetsov, V.; Lanfranchi, G.; Lamanna, G.; Latino, G.; Laycock, P.; Lazzeroni, C.; Lenti, M.; Lehmann Miotto, G.; Leonardi, E.; Lichard, P.; Litov, L.; Lollini, R.; Lomidze, D.; Lonardo, A.; Lubrano, P.; Lupi, M.; Lurkin, N.; Madigozhin, D.; Mannelli, I.; Mannocchi, G.; Mapelli, A.; Marchetto, F.; Marchevski, R.; Martellotti, S.; Massarotti, P.; Massri, K.; Maurice, E.; Medvedeva, M.; Mefodev, A.; Menichetti, E.; Migliore, E.; Minucci, E.; Mirra, M.; Misheva, M.; Molokanova, N.; Moulson, M.; Movchan, S.; Napolitano, M.; Neri, I.; Newson, F.; Norton, A.; Noy, M.; Numao, T.; Obraztsov, V.; Ostankov, A.; Padolski, S.; Page, R.; Palladino, V.; Parkinson, C.; Pedreschi, E.; Pepe, M.; Perrin-Terrin, M.; Peruzzo, L.; Petrov, P.; Petrucci, F.; Piandani, R.; Piccini, M.; Pinzino, J.; Polenkevich, I.; Pontisso, L.; Potrebenikov, Yu.; Protopopescu, D.; Raggi, M.; Romano, A.; Rubin, P.; Ruggiero, G.; Ryjov, V.; Salamon, A.; Santoni, C.; Saracino, G.; Sargeni, F.; Semenov, V.; Sergi, A.; Shaikhiev, A.; Shkarovskiy, S.; Soldi, D.; Sougonyaev, V.; Sozzi, M.; Spadaro, T.; Spinella, F.; Sturgess, A.; Swallow, J.; Trilov, S.; Valente, P.; Velghe, B.; Venditti, S.; Vicini, P.; Volpe, R.; Vormstein, M.; Wahl, H.; Wanke, R.; Wrona, B.; Yushchenko, O.; Zamkovsky, M.; Zinchenko, A.</p> <p>2018-05-01</p> <p>In this paper we present new results on upper limits for the search of Heavy Neutral Leptons (HNL) with data collected by <span class="hlt">NA</span>48/<span class="hlt">2</span> (2003-2004), <span class="hlt">NA</span>62-RK (2007) and <span class="hlt">NA</span>62 (2015) CERN experiments. The data collected with different trigger configuration allow to search for both long and short living heavy neutrinos in the mass range below the kaon mass. In addition the status of the search for K+ → π+v<overline>v</overline> with the <span class="hlt">NA</span>62 detector will be briefly presented.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28813588','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28813588"><span>Insight to the Thermal Decomposition and Hydrogen Desorption Behaviors of <span class="hlt">Na</span>NH<span class="hlt">2</span>-<span class="hlt">Na</span>BH4 Hydrogen Storage Composite.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Pei, Ziwei; Bai, Ying; Wang, Yue; Wu, Feng; Wu, Chuan</p> <p>2017-09-20</p> <p>The lightweight compound material <span class="hlt">Na</span>NH <span class="hlt">2</span> -<span class="hlt">Na</span>BH 4 is regarded as a promising hydrogen storage composite due to the high hydrogen density. Mechanical ball milling was employed to synthesize the composite <span class="hlt">Na</span>NH <span class="hlt">2</span> -<span class="hlt">Na</span>BH 4 (<span class="hlt">2</span>/1 molar ratio), and the samples were investigated utilizing thermogravimetric-differential thermal analysis-mass spectroscopy (TG-DTA-MS), X-ray diffraction (XRD), and Fourier transform infrared spectroscopy (FTIR) analyses. The full-spectrum test (range of the ratio of mass to charge: 0-200) shows that the released gaseous species contain H <span class="hlt">2</span> , NH 3 , B <span class="hlt">2</span> H 6 , and N <span class="hlt">2</span> in the heating process from room temperature to 400 °C, and possibly the impurity gas B 6 H 12 also exists. The TG/DTA analyses show that the composite <span class="hlt">Na</span>NH <span class="hlt">2</span> -<span class="hlt">Na</span>BH 4 (<span class="hlt">2</span>/1 molar ratio) is conductive to generate hydrogen so that the dehydrogenation process can be finished before 400 °C. Moreover, the thermal decomposition process from 200 to 400 °C involves two-step dehydrogenation reactions: (1) <span class="hlt">Na</span> 3 (NH <span class="hlt">2</span> ) <span class="hlt">2</span> BH 4 hydride decomposes into <span class="hlt">Na</span> 3 BN <span class="hlt">2</span> and H <span class="hlt">2</span> (200-350 °C); (<span class="hlt">2</span>) remaining <span class="hlt">Na</span> 3 (NH <span class="hlt">2</span> ) <span class="hlt">2</span> BH 4 reacts with <span class="hlt">Na</span>BH 4 and <span class="hlt">Na</span> 3 BN <span class="hlt">2</span> , generating <span class="hlt">Na</span>, BN, NH 3 , N <span class="hlt">2</span> , and H <span class="hlt">2</span> (350-400 °C). The better mechanism understanding of the thermal decomposition pathway lays a foundation for tailoring the hydrogen storage performance of the composite complex hydrides system.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/1339555-synthesis-ternary-acetylides-tellurium-li-tec-na-tec','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/1339555-synthesis-ternary-acetylides-tellurium-li-tec-na-tec"><span>The synthesis of ternary acetylides with tellurium: Li <span class="hlt">2</span> TeC <span class="hlt">2</span> and <span class="hlt">Na</span> <span class="hlt">2</span> TeC <span class="hlt">2</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>Németh, Károly; Unni, Aditya K.; Kalnmals, Christopher</p> <p></p> <p>The synthesis of ternary acetylides Li <span class="hlt">2</span>TeC <span class="hlt">2</span> and <span class="hlt">Na</span> <span class="hlt">2</span>TeC <span class="hlt">2</span> is presented as the first example of ternary acetylides with metalloid elements instead of transition metals. The synthesis was carried out by the direct reaction of the corresponding bialkali acetylides with tellurium powder in liquid ammonia. Alternatively, the synthesis of <span class="hlt">Na</span> <span class="hlt">2</span>TeC <span class="hlt">2</span> was also carried out by the direct reaction of tellurium powder and two equivalents of <span class="hlt">Na</span>C <span class="hlt">2</span>H in liquid ammonia leading to <span class="hlt">Na</span> <span class="hlt">2</span>TeC <span class="hlt">2</span> and acetylene gas through an equilibrium containing the assumed <span class="hlt">Na</span>TeC <span class="hlt">2</span>H molecules besides the reactants and the products. The resultingmore » disordered crystalline materials were characterized by X-ray diffraction and Raman spectroscopy. Implications of these new syntheses on the synthesis of other ternary acetylides with metalloid elements and transition metals are also discussed.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29266642','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29266642"><span>Vacancy-Controlled <span class="hlt">Na</span>+ Superion Conduction in <span class="hlt">Na</span>11 Sn<span class="hlt">2</span> PS12.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Duchardt, Marc; Ruschewitz, Uwe; Adams, Stefan; Dehnen, Stefanie; Roling, Bernhard</p> <p>2018-01-26</p> <p>Highly conductive solid electrolytes are crucial to the development of efficient all-solid-state batteries. Meanwhile, the ion conductivities of lithium solid electrolytes match those of liquid electrolytes used in commercial Li + ion batteries. However, concerns about the future availability and the price of lithium made <span class="hlt">Na</span> + ion conductors come into the spotlight in recent years. Here we present the superionic conductor <span class="hlt">Na</span> 11 Sn <span class="hlt">2</span> PS 12 , which possesses a room temperature <span class="hlt">Na</span> + conductivity close to 4 mS cm -1 , thus the highest value known to date for sulfide-based solids. Structure determination based on synchrotron X-ray powder diffraction data proves the existence of <span class="hlt">Na</span> + vacancies. As confirmed by bond valence site energy calculations, the vacancies interconnect ion migration pathways in a 3D manner, hence enabling high <span class="hlt">Na</span> + conductivity. The results indicate that sodium electrolytes are about to equal the performance of their lithium counterparts. © 2018 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018PCM...tmp..219P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PCM...tmp..219P"><span>The system <span class="hlt">Na</span><span class="hlt">2</span>CO3-CaCO3 at 3 GPa</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Podborodnikov, Ivan V.; Shatskiy, Anton; Arefiev, Anton V.; Rashchenko, Sergey V.; Chanyshev, Artem D.; Litasov, Konstantin D.</p> <p>2018-04-01</p> <p>It was suggested that alkali-alkaline earth carbonates may have a substantial role in petrological processes relevant to metasomatism and melting of the Earth's mantle. Because natrite, <span class="hlt">Na</span><span class="hlt">2</span>CO3, <span class="hlt">Na</span>-Ca carbonate (shortite and/or nyerereite), and calcite, CaCO3, have been recently reported from xenoliths of shallow mantle (110-115 km) origin, we performed experiments on phase relations in the system <span class="hlt">Na</span><span class="hlt">2</span>CO3-CaCO3 at 3 GPa and 800-1300 °C. We found that the system has one intermediate compound, <span class="hlt">Na</span><span class="hlt">2</span>Ca3(CO3)4, at 800 °C, and two intermediate compounds, <span class="hlt">Na</span><span class="hlt">2</span>Ca(CO3)<span class="hlt">2</span> and <span class="hlt">Na</span><span class="hlt">2</span>Ca3(CO3)4, at 850 °C. CaCO3 crystals recovered from experiments at 950 and 1000 °C are aragonite and calcite, respectively. Maximum solid solution of CaCO3 in <span class="hlt">Na</span><span class="hlt">2</span>CO3 is 20 mol% at 850 °C. The <span class="hlt">Na-carbonate-Na</span><span class="hlt">2</span>Ca(CO3)<span class="hlt">2</span> eutectic locates near 860 °C and 56 mol% <span class="hlt">Na</span><span class="hlt">2</span>CO3. <span class="hlt">Na</span><span class="hlt">2</span>Ca(CO3)<span class="hlt">2</span> melts incongruently near 880 °C to produce <span class="hlt">Na</span><span class="hlt">2</span>Ca3(CO3)4 and a liquid containing about 51 mol% <span class="hlt">Na</span><span class="hlt">2</span>CO3. <span class="hlt">Na</span><span class="hlt">2</span>Ca3(CO3)4 disappears above 1000 °C via incongruent melting to calcite and a liquid containing about 43 mol% <span class="hlt">Na</span><span class="hlt">2</span>CO3. At 1050 °C, the liquid, coexisting with <span class="hlt">Na</span>-carbonate, contains 87 mol% <span class="hlt">Na</span><span class="hlt">2</span>CO3. <span class="hlt">Na</span>-carbonate remains solid up to 1150 °C and melts at 1200 °C. The <span class="hlt">Na</span><span class="hlt">2</span>CO3 content in the liquid coexisting with calcite decreases to 15 mol% as temperature increases to 1300 °C. Considering the present and previous data, a range of the intermediate compounds on the liquidus of the <span class="hlt">Na</span><span class="hlt">2</span>CO3-CaCO3 join changes as pressure increases in the following sequence: <span class="hlt">Na</span><span class="hlt">2</span>Ca(CO3)<span class="hlt">2</span> (0.1 GPa) → <span class="hlt">Na</span><span class="hlt">2</span>Ca(CO3)<span class="hlt">2</span>, <span class="hlt">Na</span><span class="hlt">2</span>Ca3(CO3)4 (3 GPa) → <span class="hlt">Na</span>4Ca(CO3)3, <span class="hlt">Na</span><span class="hlt">2</span>Ca3(CO3)4 (6 GPa). Thus, the <span class="hlt">Na</span><span class="hlt">2</span>Ca(CO3)<span class="hlt">2</span> nyerereite stability field extends to the shallow mantle pressures. Consequently, findings of nyerereite among daughter phases in the melt inclusions in olivine from the sheared garnet peridotites are consistent with their mantle origin.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19830061417&hterms=K2&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3DK2','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19830061417&hterms=K2&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3DK2"><span>Electron affinities of the alkali dimers - <span class="hlt">Na</span><span class="hlt">2</span>, K<span class="hlt">2</span>, and Rb<span class="hlt">2</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Partridge, H.; Dixon, D. A.; Walch, S. P.; Bauschlicher, C. W., Jr.; Gole, J. L.</p> <p>1983-01-01</p> <p>Ab initio calculations on the ground states of the alkali dimers, <span class="hlt">Na</span><span class="hlt">2</span>, K<span class="hlt">2</span>, and Rb<span class="hlt">2</span>, and their anions are reported. The calculations employ large Gaussian basis sets and account for nearly all of the valence correlation energy. The calculated atomic electron affinities are within 0.02 eV of experiment and the calculated adiabatic electron affinities for <span class="hlt">Na</span><span class="hlt">2</span>, K<span class="hlt">2</span>, and Rb<span class="hlt">2</span> are, respectively, 0.470, 0.512, and 0.513 eV.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018musr.confa1018U','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018musr.confa1018U"><span><span class="hlt">Na</span> Diffusion in Quasi One-Dimensional Ion Conductor <span class="hlt">Na</span>Mn<span class="hlt">2</span>O4 Observed by μ+SR</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Umegaki, Izumi; Nozaki, Hiroshi; Harada, Masashi; Månsson, Martin; Sakurai, Hiroya; Kawasaki, Ikuto; Watanabe, Isao; Sugiyama, Jun</p> <p></p> <p>A quasi one-dimensional (1D) compound, <span class="hlt">Na</span>Mn<span class="hlt">2</span>O4, in which Mn<span class="hlt">2</span>O4 zigzag chains form a 1D channel along the b-axis and <span class="hlt">Na</span> ions locate at the center of the channel, is thought to be a good <span class="hlt">Na</span> ionic conductor. In order to study <span class="hlt">Na</span>-ion diffusion, we have measured μ+SR spectra using a powder sample in the temperature range between 100 and 500 K. A diffusive behavior was clearly observed above 325 K. Assuming a thermal activate process for jump diffusion of <span class="hlt">Na</span>-ion between two nearest neighboring sites, a self diffusion coefficient of <span class="hlt">Na</span> ion (DNa) and its activation energy (Ea) were estimated as DNa = (3.1 ± 0.<span class="hlt">2</span>) × 10 - 11 cm<span class="hlt">2</span>/s at 350 K and Ea = 180(9) meV.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29282869','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29282869"><span>Emission analysis of Tb3+ -and Sm3+ -ion-doped (Li<span class="hlt">2</span> O/<span class="hlt">Na</span><span class="hlt">2</span> O/K<span class="hlt">2</span> O) and (Li<span class="hlt">2</span> O + <span class="hlt">Na</span><span class="hlt">2</span> O/Li<span class="hlt">2</span> O + K<span class="hlt">2</span> O/K<span class="hlt">2</span> O + <span class="hlt">Na</span><span class="hlt">2</span> O)-modified borosilicate glasses.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Naveen Kumar Reddy, B; Sailaja, S; Thyagarajan, K; Jho, Young Dahl; Sudhakar Reddy, B</p> <p>2018-05-01</p> <p>Four series of borosilicate glasses modified by alkali oxides and doped with Tb 3+ and Sm 3+ ions were prepared using the conventional melt quenching technique, with the chemical composition 74.5B <span class="hlt">2</span> O 3 + 10SiO <span class="hlt">2</span> + 5MgO + R + 0.5(Tb <span class="hlt">2</span> O 3 /Sm <span class="hlt">2</span> O 3 ) [where R = 10(Li <span class="hlt">2</span> O /<span class="hlt">Na</span> <span class="hlt">2</span> O/K <span class="hlt">2</span> O) for series A and C, and R = 5(Li <span class="hlt">2</span> O + <span class="hlt">Na</span> <span class="hlt">2</span> O/Li <span class="hlt">2</span> O + K <span class="hlt">2</span> O/K <span class="hlt">2</span> O + <span class="hlt">Na</span> <span class="hlt">2</span> O) for series B and D]. The X-ray diffraction (XRD) patterns of all the prepared glasses indicate their amorphous nature. The spectroscopic properties of the prepared glasses were studied by optical absorption analysis, photoluminescence excitation (PLE) and photoluminescence (PL) analysis. A green emission corresponding to the 5 D 4 → 7 F 5 (543 nm) transition of the Tb 3+ ions was registered under excitation at 379 nm for series A and B glasses. The emission spectra of the Sm 3+ ions with the series C and D glasses showed strong reddish-orange emission at 600 nm ( 4 G 5/<span class="hlt">2</span> → 6 H 7/<span class="hlt">2</span> ) with an excitation wavelength λ exci = 404 nm ( 6 H 5/<span class="hlt">2</span> → 4 F 7/<span class="hlt">2</span> ). Furthermore, the change in the luminescence intensity with the addition of an alkali oxide and combinations of these alkali oxides to borosilicate glasses doped with Tb 3+ and Sm 3+ ions was studied to optimize the potential alkali-oxide-modified borosilicate glass. Copyright © 2017 John Wiley & Sons, Ltd.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://eric.ed.gov/?q=Raise+AND+iq&pg=4&id=ED283870','ERIC'); return false;" href="https://eric.ed.gov/?q=Raise+AND+iq&pg=4&id=ED283870"><span>An Empirical Assessment of Selected Software Purported to Raise SAT Scores Significantly When Utilized With Short-Term <span class="hlt">CAI</span> on the Microcomputer.</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>Davis, Wesley D.</p> <p></p> <p>This study evaluated Krell's 1981-82 Scholastic Aptitude Test (SAT) preparatory series software purported to raise students' scores substantially after only a short term of computer-assisted instruction (<span class="hlt">CAI</span>). Forty-eight college-bound juniors from Escambia County (Florida) were assigned to experimental and control groups. A two-phased pre- and…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25927621','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25927621"><span>Photoemission study of the electronic structure and charge density waves of <span class="hlt">Na</span><span class="hlt">2</span>Ti<span class="hlt">2</span>Sb<span class="hlt">2</span>O.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Tan, S Y; Jiang, J; Ye, Z R; Niu, X H; Song, Y; Zhang, C L; Dai, P C; Xie, B P; Lai, X C; Feng, D L</p> <p>2015-04-30</p> <p>The electronic structure of <span class="hlt">Na</span><span class="hlt">2</span>Ti<span class="hlt">2</span>Sb<span class="hlt">2</span>O single crystal is studied by photon energy and polarization dependent angle-resolved photoemission spectroscopy (ARPES). The obtained band structure and Fermi surface agree well with the band structure calculation of <span class="hlt">Na</span><span class="hlt">2</span>Ti<span class="hlt">2</span>Sb<span class="hlt">2</span>O in the non-magnetic state, which indicates that there is no magnetic order in <span class="hlt">Na</span><span class="hlt">2</span>Ti<span class="hlt">2</span>Sb<span class="hlt">2</span>O and the electronic correlation is weak. Polarization dependent ARPES results suggest the multi-band and multi-orbital nature of <span class="hlt">Na</span><span class="hlt">2</span>Ti<span class="hlt">2</span>Sb<span class="hlt">2</span>O. Photon energy dependent ARPES results suggest that the electronic structure of <span class="hlt">Na</span><span class="hlt">2</span>Ti<span class="hlt">2</span>Sb<span class="hlt">2</span>O is rather two-dimensional. Moreover, we find a density wave energy gap forms below the transition temperature and reaches 65 meV at 7 K, indicating that <span class="hlt">Na</span><span class="hlt">2</span>Ti<span class="hlt">2</span>Sb<span class="hlt">2</span>O is likely a weakly correlated CDW material in the strong electron-phonon interaction regime.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4537937','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4537937"><span>Analysis of cardiovascular responses to the H<span class="hlt">2</span>S donors <span class="hlt">Na</span><span class="hlt">2</span>S and <span class="hlt">Na</span>HS in the rat</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Yoo, Daniel; Jupiter, Ryan C.; Pankey, Edward A.; Reddy, Vishwaradh G.; Edward, Justin A.; Swan, Kevin W.; Peak, Taylor C.; Mostany, Ricardo</p> <p>2015-01-01</p> <p>Hydrogen sulfide (H<span class="hlt">2</span>S) is an endogenous gaseous molecule formed from L-cysteine in vascular tissue. In the present study, cardiovascular responses to the H<span class="hlt">2</span>S donors <span class="hlt">Na</span><span class="hlt">2</span>S and <span class="hlt">Na</span>HS were investigated in the anesthetized rat. The intravenous injections of <span class="hlt">Na</span><span class="hlt">2</span>S and <span class="hlt">Na</span>HS 0.03–0.5 mg/kg produced dose-related decreases in systemic arterial pressure and heart rate, and at higher doses decreases in cardiac output, pulmonary arterial pressure, and systemic vascular resistance. H<span class="hlt">2</span>S infusion studies show that decreases in systemic arterial pressure, heart rate, cardiac output, and systemic vascular resistance are well-maintained, and responses to <span class="hlt">Na</span><span class="hlt">2</span>S are reversible. Decreases in heart rate were not blocked by atropine, suggesting that the bradycardia was independent of parasympathetic activation and was mediated by an effect on the sinus node. The decreases in systemic arterial pressure were not attenuated by hexamethonium, glybenclamide, Nw-nitro-l-arginine methyl ester hydrochloride, sodium meclofenamate, ODQ, miconazole, 5-hydroxydecanoate, or tetraethylammonium, suggesting that ATP-sensitive potassium channels, nitric oxide, arachidonic acid metabolites, cyclic GMP, p450 epoxygenase metabolites, or large conductance calcium-activated potassium channels are not involved in mediating hypotensive responses to the H<span class="hlt">2</span>S donors in the rat and that responses are not centrally mediated. The present data indicate that decreases in systemic arterial pressure in response to the H<span class="hlt">2</span>S donors can be mediated by decreases in vascular resistance and cardiac output and that the donors have an effect on the sinus node independent of the parasympathetic system. The present data indicate that the mechanism of the peripherally mediated hypotensive response to the H<span class="hlt">2</span>S donors is uncertain in the intact rat. PMID:26071540</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=225291','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=225291"><span>MEDLEARN: a computer-assisted instruction (<span class="hlt">CAI</span>) program for MEDLARS.</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Eisenberg, L J; Standing, R A; Tidball, C S; Leiter, J</p> <p>1978-01-01</p> <p>*MEDLEARN*, a second-generation computer-assisted instruction (<span class="hlt">CAI</span>) program available (nationally) since October 1976, provides on-line training for MEDLINE, one of the National Library of Medicine's (NLM) Medical Literature Analysis and Retrieval System (MEDLARS) data base. *MEDLEARN* was developed as a joint effort between NLM and The George Washington University Medical Center. Using MEDLINE formats throughout, *MEDLEARN* combines tutorial dialogue, drill and practice, testing, and simulation. The program was designed in three tracks oriented to basic methods, advanced techniques, and new developments. Each topic is presented on two levels, permitting an alternate explanation for users encountering difficulty. *MEDLEARN*, coded in the computer language PILOT, was developed with a modular structure which promotes ease of writing and revision. A versatile control structure maximizes student control. Frequent interactions check immediate recall, general comprehension, and integration of knowledge. Two MEDLINE simulations are included, providing the student an opportunity to formulate and execute a search, have it evaluated, and then perform the search in MEDLINE. Commenting, news broadcasting, and monitoring (with permission only) capabilities are also available. Subjective field appraisals have been positive and NLM plans to expand *MEDLEARN* and produce similar programs for other data bases. PMID:342015</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li class="active"><span>11</span></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_11 --> <div id="page_12" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li class="active"><span>12</span></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="221"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3681939','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3681939"><span>Intracellular and Extracellular pH and Ca Are Bound to Control Mitosis in the Early Sea Urchin Embryo via ERK and MPF Activities</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Ciapa, Brigitte; Philippe, Laetitia</p> <p>2013-01-01</p> <p>Studies aiming to predict the impact on marine life of ocean acidification and of altered salinity have shown altered development in various species including sea urchins. We have analyzed how external <span class="hlt">Na</span>, Ca, pH and bicarbonate control the first mitotic divisions of sea urchin embryos. Intracellular free Ca (<span class="hlt">Cai</span>) and pH (pHi) and the activities of the MAP kinase ERK and of MPF regulate mitosis in various types of cells including oocytes and early embryos. We found that intracellular acidification of fertilized eggs by <span class="hlt">Na</span>-acetate induces a huge activation of ERK at time of mitosis. This also stops the cell cycle and leads to cell death, which can be bypassed by treatment with the MEK inhibitor U0126. Similar intracellular acidification induced in external medium containing low sodium or 5-(N-Methyl-N-isobutyl) amiloride, an inhibitor of the <span class="hlt">Na</span>+/H+ exchanger, also stops the cell cycle and leads to cell death. In that case, an increase in <span class="hlt">Cai</span> and in the phosphorylation of tyr-cdc<span class="hlt">2</span> occurs during mitosis, modifications that depend on external Ca. Our results indicate that the levels of pHi and <span class="hlt">Cai</span> determine accurate levels of Ptyr-Cdc<span class="hlt">2</span> and P-ERK capable of ensuring progression through the first mitotic cycles. These intracellular parameters rely on external Ca, <span class="hlt">Na</span> and bicarbonate, alterations of which during climate changes could act synergistically to perturb the early marine life. PMID:23785474</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23785474','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23785474"><span>Intracellular and extracellular pH and Ca are bound to control mitosis in the early sea urchin embryo via ERK and MPF activities.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Ciapa, Brigitte; Philippe, Laetitia</p> <p>2013-01-01</p> <p>Studies aiming to predict the impact on marine life of ocean acidification and of altered salinity have shown altered development in various species including sea urchins. We have analyzed how external <span class="hlt">Na</span>, Ca, pH and bicarbonate control the first mitotic divisions of sea urchin embryos. Intracellular free Ca (<span class="hlt">Cai</span>) and pH (pHi) and the activities of the MAP kinase ERK and of MPF regulate mitosis in various types of cells including oocytes and early embryos. We found that intracellular acidification of fertilized eggs by <span class="hlt">Na</span>-acetate induces a huge activation of ERK at time of mitosis. This also stops the cell cycle and leads to cell death, which can be bypassed by treatment with the MEK inhibitor U0126. Similar intracellular acidification induced in external medium containing low sodium or 5-(N-Methyl-N-isobutyl) amiloride, an inhibitor of the <span class="hlt">Na</span>(+)/H(+) exchanger, also stops the cell cycle and leads to cell death. In that case, an increase in <span class="hlt">Cai</span> and in the phosphorylation of tyr-cdc<span class="hlt">2</span> occurs during mitosis, modifications that depend on external Ca. Our results indicate that the levels of pHi and <span class="hlt">Cai</span> determine accurate levels of Ptyr-Cdc<span class="hlt">2</span> and P-ERK capable of ensuring progression through the first mitotic cycles. These intracellular parameters rely on external Ca, <span class="hlt">Na</span> and bicarbonate, alterations of which during climate changes could act synergistically to perturb the early marine life.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1997MolPh..91..917M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1997MolPh..91..917M"><span>Electronic structure and molecular dynamics of <span class="hlt">Na</span><span class="hlt">2</span>Li</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Malcolm, Nathaniel O. J.; McDouall, Joseph J. W.</p> <p></p> <p>Following the first report (Mile, B., Sillman, P. D., Yacob, A. R. and Howard, J. A., 1996, J. chem. Soc. Dalton Trans , 653) of the EPR spectrum of the mixed alkali-metal trimer <span class="hlt">Na</span><span class="hlt">2</span>Li a detailed study has been made of the electronic structure and structural dynamics of this species. Two isomeric forms have been found: one of the type, <span class="hlt">Na-Li-Na</span>, of C , symmetry and another, Li-<span class="hlt">Na-Na</span>, of C symmetry. Also, there are two linear saddle points which correspond to 'inversion' transition structures, and a saddle point of C symmetry which connects the two minima. A molecular dynamics investigation of these species shows that, at the temperature of the reported experiments (170 K), the C minimum is not 'static', but undergoes quite rapid inversion. At higher temperatures the C minimum converts to the C form, but by a mechanism very different from that suggested by minimum energy path considerations. <span class="hlt">2</span> <span class="hlt">2</span>v s s <span class="hlt">2</span>v <span class="hlt">2</span>v s</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23295625','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23295625"><span>Fluoride gastrointestinal absorption from <span class="hlt">Na</span><span class="hlt">2</span>FPO3/CaCO3- and <span class="hlt">Na</span>F/SiO<span class="hlt">2</span>-based toothpastes.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Falcão, A; Tenuta, L M A; Cury, J A</p> <p>2013-01-01</p> <p>Depending on toothpaste formulation, part of the fluoride is insoluble and would not be totally absorbable in the gastrointestinal tract, thus changing dental fluorosis risk estimation. This hypothesis was tested with formulations with either all fluoride in a soluble form (<span class="hlt">Na</span>F/SiO<span class="hlt">2</span>-based toothpaste, 1,100 µg F/g as labeled, 1,129.7 ± 49.4 µg F/g soluble fluoride as analyzed) or with around 20% of insoluble fluoride (<span class="hlt">Na</span><span class="hlt">2</span>FPO3/CaCO3-based toothpaste, 1,450 µg F/g as labeled, 1,122.4 ± 76.4 µg F/g soluble fluoride as analyzed). Toothpastes were evaluated either fresh or after accelerated aging, which increased insoluble fluoride to 40% in the <span class="hlt">Na</span><span class="hlt">2</span>FPO3/CaCO3-based toothpaste. In a blind, crossover clinical trial conducted in five legs, 20 adult volunteers ingested 49.5 µg of total fluoride/kg body weight from each formulation or purified water (control). Whole saliva and urine were collected as bioavailability indicators, and pharmacokinetics parameters calculated showed significantly (p < 0.05) lower fluoride bioavailability for <span class="hlt">Na</span><span class="hlt">2</span>FPO3/CaCO3 toothpaste, which was reduced further after aging. A significant correlation between the amount of soluble fluoride ingested, but not total fluoride, and fluoride bioavailability was found (r = 0.57, p < 0.0001). The findings suggest that the estimated fluorosis risk as a result of ingestion of <span class="hlt">Na</span><span class="hlt">2</span>FPO3/CaCO3-based toothpastes should be calculated based on the toothpaste's soluble rather than total fluoride concentration. Copyright © 2012 S. Karger AG, Basel.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/11718362','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/11718362"><span>Enhancement of the sulfur capture capacity of limestones by the addition of <span class="hlt">Na</span><span class="hlt">2</span>CO3 and <span class="hlt">Na</span>Cl.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Laursen, K; Grace, J R; Lim, C J</p> <p>2001-11-01</p> <p>The ability of <span class="hlt">Na</span><span class="hlt">2</span>CO3 and <span class="hlt">Na</span>Cl to enhance the sulfur capture capacity of three limestones was evaluated via fixed-bed calcination and sulfation experiments. The tested limestones represent three different sulfation morphologies: unreacted-core, network, and uniformly sulfated. Treatment with aqueous or powdered <span class="hlt">Na</span><span class="hlt">2</span>CO3 significantly increased the Ca-utilization for two stones which normally sulfate in an unreacted-core pattern (20% to 45%) and network pattern (33% to 49%). The increase was lower for the uniformly sulfated stone (44% to 48%). <span class="hlt">Na</span><span class="hlt">2</span>CO3 treatment increased the number of macropores leading to uniform sulfation of all particles, nearly eliminating the normal strong dependence of utilization on limestone type and particle size. The effect of <span class="hlt">Na</span><span class="hlt">2</span>CO3 is believed to be associated with formation of a eutectic melt which enhances ionic diffusion and accelerates molecular rearrangement of the CaO. Treatment with aqueous <span class="hlt">Na</span>Cl solution caused a decrease in utilization, probably due to formation of large grains and plugging of pores caused by formation of a large amount of eutectic melt. The effect of <span class="hlt">Na</span><span class="hlt">2</span>CO3 is less sensitive than that of <span class="hlt">Na</span>Cl to the amount added and the combustion environment (temperature and gas composition). In addition, <span class="hlt">Na</span><span class="hlt">2</span>CO3 neither promotes corrosion nor forms chlorinated byproducts, which are main concerns associated with <span class="hlt">Na</span>Cl. Thus, <span class="hlt">Na</span><span class="hlt">2</span>CO3 appears to have significant advantages over <span class="hlt">Na</span>Cl for enhancement of limestone sulfur capture capacity in fluidized-bed combustors.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1980MTB....11..607N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1980MTB....11..607N"><span>Thermodynamic properties of <span class="hlt">Na</span><span class="hlt">2</span>O-SiO<span class="hlt">2</span>-CaO melts at 1000 to 1100 °C</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Neudorf, D. A.; Elliott, J. F.</p> <p>1980-12-01</p> <p>The thermodynamic properties of <span class="hlt">Na</span><span class="hlt">2</span>O-SiO<span class="hlt">2</span> and <span class="hlt">Na</span><span class="hlt">2</span>O-SiO<span class="hlt">2</span>-CaO melts have been measured using the galvanic cellbegin{array}{*{20}c} {O_<span class="hlt">2</span> (g), (<span class="hlt">Na</span>_<span class="hlt">2</span> O), Pt} \\ {<span class="hlt">Na</span>_<span class="hlt">2</span> O - WO_3 liq} \\ left| begin{gathered} <span class="hlt">Na</span>^ + \\ β - alumina \\ right| begin{array}{*{20}c} {Pt,(<span class="hlt">Na</span>_<span class="hlt">2</span> O), O_<span class="hlt">2</span> (g)} \\ {<span class="hlt">Na</span>_<span class="hlt">2</span> O - SiO_<span class="hlt">2</span> - CaO liq} \\ Activities of <span class="hlt">Na</span><span class="hlt">2</span>O were calculated from the reversible emf of the cell. This is possible because the activity of <span class="hlt">Na</span><span class="hlt">2</span>O in the <span class="hlt">Na</span><span class="hlt">2</span>O-WO3 liquid is known from previous work. Data for the binary <span class="hlt">Na</span><span class="hlt">2</span>O-SiO<span class="hlt">2</span> system were obtained between 1000 and 1100 °C and for compositions ranging from 25 wt pct to 40 wt pct <span class="hlt">Na</span><span class="hlt">2</span>O. At 1050 °C, Loga_{<span class="hlt">Na</span>_<span class="hlt">2</span> O} varied from approximately 10.<span class="hlt">2</span> at 25 wt pct <span class="hlt">Na</span><span class="hlt">2</span>O to approximately -8.3 at 40 wt pct <span class="hlt">Na</span><span class="hlt">2</span>O, the dependence with respect to composition being nearly linear. The Gibbs-Duhem equation was used to calculate the activities of SiO<span class="hlt">2</span>(s), and the integral mixing properties, G M, HM, and S M, were derived. At the di-silicate composition, G M = -83 kJ/mol, H M = -41 kJ mol and S M = 33 J/mol K at 1000 °C. (Standard states are pure, liquid <span class="hlt">Na</span><span class="hlt">2</span>O and pure, solid tridymite.) The activity data are interpreted in terms of the polymeric nature of silicate melts. Activities of <span class="hlt">Na</span><span class="hlt">2</span>O in the <span class="hlt">Na</span><span class="hlt">2</span>O-CaO-SiO<span class="hlt">2</span> system were measured for the 25, 30 and 35 wt pct <span class="hlt">Na</span><span class="hlt">2</span>O binary compositions with up to 10 wt pct CaO added. The addition of CaO caused an increase in the activity of <span class="hlt">Na</span><span class="hlt">2</span>O at constantN_{<span class="hlt">Na</span>_<span class="hlt">2</span> O} /N_{SiO_<span class="hlt">2</span> } . The experimental data agree well with the behavior predicted by Richardson’s ternary mixing model.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-KSC-20170222-PH_ELA01_0002.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-KSC-20170222-PH_ELA01_0002.html"><span>CAPE-<span class="hlt">2</span> Cubesat - ELa<span class="hlt">Na</span> IV</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2016-07-25</p> <p>CAPE-<span class="hlt">2</span>: Cajun Advanced Picosatellite Experiment – ELa<span class="hlt">Na</span> IV CAPE-<span class="hlt">2</span> was developed by students from the University of Louisiana Lafayette to engage, inspire and educate K-12 students to encourage them to pursue STEM careers. The secondary focus is the technology demonstration of deployed solar panels to support the following payloads: text to speech, voice repeater, tweeting, email, file transfer and data collection from buoys. Launched by NASA’s CubeSat Launch Initiative on the ELa<span class="hlt">Na</span> IV mission as an auxiliary payload aboard the U.S. Air Force-led Operationally Responsive Space (ORS-3) Mission on November 19, 2013.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28374959','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28374959"><span>Hierarchical Porous Carbon Spheres for High-Performance <span class="hlt">Na</span>-O<span class="hlt">2</span> Batteries.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Sun, Bing; Kretschmer, Katja; Xie, Xiuqiang; Munroe, Paul; Peng, Zhangquan; Wang, Guoxiu</p> <p>2017-12-01</p> <p>As a new family member of room-temperature aprotic metal-O <span class="hlt">2</span> batteries, <span class="hlt">Na</span>-O <span class="hlt">2</span> batteries, are attracting growing attention because of their relatively high theoretical specific energy and particularly their uncompromised round-trip efficiency. Here, a hierarchical porous carbon sphere (PCS) electrode that has outstanding properties to realize <span class="hlt">Na</span>-O <span class="hlt">2</span> batteries with excellent electrochemical performances is reported. The controlled porosity of the PCS electrode, with macropores formed between PCSs and nanopores inside each PCS, enables effective formation/decomposition of <span class="hlt">Na</span>O <span class="hlt">2</span> by facilitating the electrolyte impregnation and oxygen diffusion to the inner part of the oxygen electrode. In addition, the discharge product of <span class="hlt">Na</span>O <span class="hlt">2</span> is deposited on the surface of individual PCSs with an unusual conformal film-like morphology, which can be more easily decomposed than the commonly observed microsized <span class="hlt">Na</span>O <span class="hlt">2</span> cubes in <span class="hlt">Na</span>-O <span class="hlt">2</span> batteries. A combination of coulometry, X-ray diffraction, and in situ differential electrochemical mass spectrometry provides compelling evidence that the operation of the PCS-based <span class="hlt">Na</span>-O <span class="hlt">2</span> battery is underpinned by the formation and decomposition of <span class="hlt">Na</span>O <span class="hlt">2</span> . This work demonstrates that employing nanostructured carbon materials to control the porosity, pore-size distribution of the oxygen electrodes, and the morphology of the discharged <span class="hlt">Na</span>O <span class="hlt">2</span> is a promising strategy to develop high-performance <span class="hlt">Na</span>-O <span class="hlt">2</span> batteries. © 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013JSSCh.207...21S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013JSSCh.207...21S"><span><span class="hlt">Na</span>8Au9.8(4)Ga7.<span class="hlt">2</span> and <span class="hlt">Na</span>17Au5.87(<span class="hlt">2</span>)Ga46.63: The diversity of pseudo 5-fold symmetries in the <span class="hlt">Na</span>-Au-Ga system</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Smetana, Volodymyr; Corbett, John D.; Miller, Gordon J.</p> <p>2013-11-01</p> <p>The <span class="hlt">Na</span>-rich part (~30% <span class="hlt">Na</span>) of the <span class="hlt">Na</span>-Au-Ga system between <span class="hlt">Na</span>Au<span class="hlt">2</span>, <span class="hlt">Na</span>Ga4, and <span class="hlt">Na</span>22Ga39 has been found to contain the ternary phases <span class="hlt">Na</span>8Au9.8(4)Ga7.<span class="hlt">2</span> (I) and <span class="hlt">Na</span>17Au5.87(<span class="hlt">2</span>)Ga46.63 (II), according to the results of single crystal X-ray diffraction measurements. I is orthorhombic, Cmcm, a=5.3040(1), b=24.519(5), c=14.573(3) Å, and contains a network of clusters with local 5-fold symmetry along the a-axis. Such clusters are frequent building units in decagonal quasicrystals and their approximants. II is rhombohedral, R3¯m, a=16.325(<span class="hlt">2</span>), c=35.242(7) Å, and contains building blocks that are structurally identical to the Bergman-type clusters as well as fused icosahedral units known with active metals, triels and late transition elements. II also contains a polycationic network with elements of the clathrate V type structure. Tight-binding electronic structure calculations using linear muffin-tin-orbital (LMTO) methods on idealized models of I and II indicate that both compounds are metallic with evident pseudogaps at the corresponding Fermi levels. The overall Hamilton bond populations are generally dominated by Au-Ga and Au-Au bonds in I and by Ga-Ga bonds in II; moreover, the <span class="hlt">Na</span>-Au and <span class="hlt">Na</span>-Ga contributions in I are unexpectedly large, ~20% of the total. A similar involvement of sodium in covalent bonding has also been found in the electron-richer i-<span class="hlt">Na</span>13Au12Ga15 quasicrystal approximant.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29352355','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29352355"><span>Physiological responses to salt stress of salt-adapted and directly salt (<span class="hlt">Na</span>Cl and <span class="hlt">NaCl+Na</span><span class="hlt">2</span>SO4 mixture)-stressed cyanobacterium Anabaena fertilissima.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Swapnil, Prashant; Rai, Ashwani K</p> <p>2018-05-01</p> <p>Soil salinity in nature is generally mixed type; however, most of the studies on salt toxicity are performed with <span class="hlt">Na</span>Cl and little is known about sulfur type of salinity (<span class="hlt">Na</span> <span class="hlt">2</span> SO 4 ). Present study discerns the physiologic mechanisms responsible for salt tolerance in salt-adapted Anabaena fertilissima, and responses of directly stressed parent cells to <span class="hlt">Na</span>Cl and <span class="hlt">NaCl+Na</span> <span class="hlt">2</span> SO 4 mixture. <span class="hlt">Na</span>Cl at 500 mM was lethal to the cyanobacterium, whereas salt-adapted cells grew luxuriantly. Salinity impaired gross photosynthesis, electron transport activities, and respiration in parent cells, but not in the salt-adapted cells, except a marginal increase in PSI activity. Despite higher <span class="hlt">Na</span> + concentration in the salt mixture, equimolar <span class="hlt">Na</span>Cl appeared more inhibitive to growth. Sucrose and trehalose content and antioxidant activities were maximal in 250 mM <span class="hlt">Na</span>Cl-treated cells, followed by salt mixture and was almost identical in salt-adapted (exposed to 500 mm <span class="hlt">Na</span>Cl) and control cells, except a marginal increase in ascorbate peroxidase activity and an additional fourth superoxide dismutase isoform. Catalase isoform of 63 kDa was induced only in salt-stressed cells. Salinity increased the uptake of intracellular <span class="hlt">Na</span> + and Ca <span class="hlt">2</span>+ and leakage of K + in parent cells, while cation level in salt-adapted cells was comparable to control. Though there was differential increase in intracellular Ca <span class="hlt">2</span>+ under different salt treatments, ratio of Ca <span class="hlt">2</span>+ /<span class="hlt">Na</span> + remained the same. It is inferred that stepwise increment in the salt concentration enabled the cyanobacterium to undergo priming effect and acquire robust and efficient defense system involving the least energy.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24984488','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24984488"><span>[Effect of <span class="hlt">Na</span>HCO3 stress on uptake and transportation of <span class="hlt">Na</span>+, K+ and Ca<span class="hlt">2</span>+ in three shrub species].</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Mao, Gui-Lian; Li, Guo-Qi; Xu, Xing; Zhang, Xin-Xue</p> <p>2014-03-01</p> <p>We detected absorption and transportation of ions in the leaves of Atriplex nummularia, Atriplex canescens and Lycium barbarum under <span class="hlt">Na</span>HCO3 stress (300 mmol x L(-1)) by using atomic absorption spectrophotometry and non-invasive ion flux measurement. The results showed that leaves of the A. nummularia, A. canescens and L. barbarum exhibited a high capacity to induce the <span class="hlt">Na</span>+ accumulation when compared with that of control. The higher the concentration of <span class="hlt">Na</span>HCO3 treatment, the more <span class="hlt">Na</span>+ accumulated in the leaves of the three plants under experimental condition. L. barbarum showed a higher <span class="hlt">Na</span>+ efflux in the mesophyll cells, whereas A. nummularia and A. canescens showed a relative lower efflux. A lower K+ content and a higher <span class="hlt">Na</span>+/K+ ratio were detected in leaves of A. nummularia and L. barbarum. However, a higher K+ content and a lower <span class="hlt">Na</span>+/K+ ratio were seen in leaves of A. canescens. Due to induction of Ca<span class="hlt">2</span>+ efflux under the <span class="hlt">Na</span>HCO3 treatment, a lower Ca<span class="hlt">2</span>+ content and a higher <span class="hlt">Na</span>+/Ca<span class="hlt">2</span>+ ratio were observed in L. barbarum. On the contrary, a higher Ca<span class="hlt">2</span>+ influx was observed in A. nummularia and A. canescens. These results suggested that the three shrubs species had different <span class="hlt">Na</span>+ segmentation strategies. The accumulation of <span class="hlt">Na</span>+ inhibited Ca<span class="hlt">2</span>+ absorption in leaves of L. barbarum, while in the A. nummularia and A. canescens, Ca<span class="hlt">2</span>+ influx induced [Ca<span class="hlt">2</span>+]cyt which preserved a less-depolarized PM and then inhibited K efflux. The maintaining of cellular K+/<span class="hlt">Na</span>+ homeostasis in A. nummularia and A. canescens might be achieved by the induction of [Ca<span class="hlt">2</span>+]cyt under the <span class="hlt">Na</span>HCO3 treatment.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=185861','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=185861"><span>Urea inhibits <span class="hlt">Na</span>K<span class="hlt">2</span>Cl cotransport in human erythrocytes.</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Lim, J; Gasson, C; Kaji, D M</p> <p>1995-01-01</p> <p>We examined the effect of urea on <span class="hlt">Na</span>K<span class="hlt">2</span>Cl cotransport in human erythrocytes. In erythrocytes from nine normal subjects, the addition of 45 mM urea, a concentration commonly encountered in uremic subjects, inhibited <span class="hlt">Na</span>K<span class="hlt">2</span>Cl cotransport by 33 +/- 7%. Urea inhibited <span class="hlt">Na</span>K<span class="hlt">2</span>Cl cotransport reversibly, and in a concentration-dependent fashion with half-maximal inhibition at 63 +/- 10 mM. Acute cell shrinkage increased, and acute cell swelling decreased <span class="hlt">Na</span>K<span class="hlt">2</span>Cl cotransport in human erythrocytes. Okadaic acid (OA), a specific inhibitor of protein phosphatase 1 and <span class="hlt">2</span>A, increased <span class="hlt">Na</span>K<span class="hlt">2</span>Cl cotransport by nearly 80%, suggesting an important role for these phosphatases in the regulation of <span class="hlt">Na</span>K<span class="hlt">2</span>Cl cotransport. Urea inhibited bumetanide-sensitive K influx even when protein phosphatases were inhibited with OA, suggesting that urea acted by inhibiting a kinase. In cells subjected to shrinking and OA pretreatment, maneuvers expected to increase the net phosphorylation, urea inhibited cotransport only minimally, suggesting that urea acted by causing a net dephosphorylation of the cotransport protein, or some key regulatory protein. The finding that concentrations of urea found in uremic subjects inhibited <span class="hlt">Na</span>K<span class="hlt">2</span>Cl cotransport, a widespread transport pathway with important physiological functions, suggests that urea is not only a marker for accumulation of other uremic toxins, but may be a significant uremic toxin itself. PMID:7593597</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19990113086&qs=N%3D0%26Ntk%3DTitle%26Ntx%3Dmode%2Bmatchall%26Ntt%3DG','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19990113086&qs=N%3D0%26Ntk%3DTitle%26Ntx%3Dmode%2Bmatchall%26Ntt%3DG"><span>Heterogeneous reactions of HNO3(g) + <span class="hlt">Na</span>Cl(s) yields HCl(g) + <span class="hlt">Na</span>NO3(s) and N<span class="hlt">2</span>O5(g) + <span class="hlt">Na</span>Cl(s) yields ClNO<span class="hlt">2</span>(g) + <span class="hlt">Na</span>NO3(s)</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Leu, Ming-Taun; Timonen, Raimo S.; Keyser, Leon F.; Yung, Yuk L.</p> <p>1995-01-01</p> <p>The heterogeneous reactions of HNO3(g) + <span class="hlt">Na</span>Cl(s) yields HCl(g) + <span class="hlt">Na</span>NO3(s) (eq 1) and N<span class="hlt">2</span>O5(g) + <span class="hlt">Na</span>Cl(s) yields ClNO<span class="hlt">2</span>(g) + <span class="hlt">Na</span>NO3(S) (eq <span class="hlt">2</span>) were investigated over the temperature range 223-296 K in a flow-tube reactor coupled to a quadrupole mass spectrometer. Either a chemical ionization mass spectrometer (CIMS) or an electron-impact ionization mass spectrometer (EIMS) was used to provide suitable detection sensitivity and selectivity. In order to mimic atmospheric conditions, partial pressures of HNO3 and N<span class="hlt">2</span>O5 in the range 6 x 10(exp -8) - <span class="hlt">2</span> x 10(exp -6) Torr were used. Granule sizes and surface roughness of the solid <span class="hlt">Na</span>Cl substrates were determined by using a scanning electron microscope. For dry <span class="hlt">Na</span>Cl substrates, decay rates of HNO3 were used to obtain gamma(1) = 0.013 +/- 0.004 (1sigma) at 296 K and > 0.008 at 223 K, respectively. The error quoted is the statistical error. After all corrections were made, the overall error, including systematic error, was estimated to be about a factor of <span class="hlt">2</span>. HCl was found to be the sole gas-phase product of reaction 1. The mechanism changed from heterogeneous reaction to predominantly physical adsorption when the reactor was cooled from 296 to 223 K. For reaction <span class="hlt">2</span> using dry salts, gamma(<span class="hlt">2</span>) was found to be less than 1.0 x 10(exp -4) at both 223 and 296 K. The gas-phase reaction product was identified as ClNO<span class="hlt">2</span> in previous studies using an infrared spectrometer. An enhancement in reaction probability was observed if water was not completely removed from salt surfaces, probably due to the reaction of N<span class="hlt">2</span>O5(g) + H<span class="hlt">2</span>O(s) yields <span class="hlt">2</span>HNO3(g). Our results are compared with previous literature values obtained using different experimental techniques and conditions. The implications of the present results for the enhancement of the hydrogen chloride column density in the lower stratosphere after the El Chichon volcanic eruption and for the chemistry of HCl and HNO3 in the marine troposphere are discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19840013308&hterms=molten+salt+electrolysis&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dmolten%2Bsalt%2Belectrolysis','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19840013308&hterms=molten+salt+electrolysis&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dmolten%2Bsalt%2Belectrolysis"><span>Moderate temperature rechargeable <span class="hlt">Na</span>NiS<span class="hlt">2</span> cells</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Abraham, K. M.</p> <p>1983-01-01</p> <p>A rechargeable sodium battery of the configuration, liquid <span class="hlt">Na</span>/beta double prime -Al<span class="hlt">2</span>O3/molten <span class="hlt">Na</span>AlCl4, NiS<span class="hlt">2</span>, operating in the temperature range of 170 to 190 C, is described. This battery is capable of delivering or = to 50 W-hr/1b and 1000 deep discharge/charge cycles.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19780040992&hterms=incubation+period&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dincubation%2Bperiod','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19780040992&hterms=incubation+period&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dincubation%2Bperiod"><span>The effect of <span class="hlt">Na</span>Cl/g/ on the <span class="hlt">Na</span><span class="hlt">2</span>SO4-induced hot corrosion of NiAl</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Smeggil, J. G.; Bornstein, N. S.; Decrescente, M. A.</p> <p>1977-01-01</p> <p>Studies have been performed to examine the effect of <span class="hlt">Na</span>Cl vapor on the <span class="hlt">Na</span><span class="hlt">2</span>SO4-induced hot corrosion of the alumina former NiAl. In the incubation period associated with such hot corrosion, <span class="hlt">Na</span>Cl(g) has been shown to be effective in removing aluminum from below the protective alumina scale and redepositing it as Al<span class="hlt">2</span>O3 whiskers on the surface of the <span class="hlt">Na</span><span class="hlt">2</span>SO4-coated sample. Similar effects seen in simple oxidation are associated with isothermal rupturing of the protective alumina scale.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2003EAEJA....14093M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2003EAEJA....14093M"><span>Influence of pH and ionic strength (<span class="hlt">NaCl/Na</span><span class="hlt">2</span>SO4) on the reaction HO Cl/ClO- + NO<span class="hlt">2</span>-</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Marcellos da Rosa, M.; Zetzsch, C.</p> <p>2003-04-01</p> <p>Equilibria such as HOCl + NO_<span class="hlt">2</span>^- leftrightarrow ClNO_<span class="hlt">2</span> + OH^- and ClNO_<span class="hlt">2</span> + H_<span class="hlt">2</span>O leftrightarrow NO_3^- + <span class="hlt">2</span>H^+ + Cl^- play an important role in halogen activation in the troposphere. We studied the oxidation of NO_<span class="hlt">2</span>^- by HOCl/ClO^- in aqueous phase by stopped-flow measurements at different ionic strengths (bidestilled water, 0.1M <span class="hlt">Na</span>Cl, 1.0M <span class="hlt">Na</span>Cl and 1.0M <span class="hlt">Na</span>_<span class="hlt">2</span>SO^4) at various pH values (4.0, 5.5, 6.<span class="hlt">2</span> and 10.0) at 293K. The experiments were performed using a SX.18MV Applied Photophysics spectrophotometer, observing the exponential decay of HOCl/ClO^- at λ = 290nm between 10ms and 100s. HOCl (pK_a= 7.50) was obtained by bubbling N_<span class="hlt">2</span> with 1% Cl_<span class="hlt">2</span> through bidestilled water. The pH of the aqueous solutions of HOCl was determined by a pH meter (CG820, Schott) with a glass electrode N6180 (calibrated with standard buffer solutions at pH = 3.0, 4.0, 7.0 and 10.0), and the pH values were adjusted by dropwise addition of HClO_4 or <span class="hlt">Na</span>OH. The concentrations of HOCl (ɛHOCl (230nm) = 100M-1cm-1) ([HOCl] = 1.3mM - 10mM) and ClO- (ɛClO- (292nm) = 350 M-1cm-1) ([ClO^-] = 1.3mM - 5mM) were determined by UV spectrometry (Kontron UVIKON 860) at a resolution of <span class="hlt">2</span> nm in 1 cm cells at various pH values. The concentration range of NO_<span class="hlt">2</span>^- was between 5mM and 50mM. The following second-order rate constant kII were obtained at 293K at various pH values (in units of M-1s-1) in H_<span class="hlt">2</span>O: pH 4.0, (5.6±0.3)\\cdot 10^3; pH 5.5, (5.0±0.4)\\cdot 10^3; pH 10.0, 3.9±0.4; in 0.1M <span class="hlt">Na</span>Cl: pH 5.5, (4.3±0.4)\\cdot 10^3; pH 10.0, <span class="hlt">2</span>.6±0.4; in 1.0M <span class="hlt">Na</span>Cl: pH 5.5, (4.0±0.3); pH 10.0, 0.7±0.<span class="hlt">2</span> and in 1.0M <span class="hlt">Na</span>_<span class="hlt">2</span>SO_4: pH 5.5, (3.0±0.3)\\cdot 10^3; pH 10.0, 1.9±0.4. There is a strong effect of the pH on the reaction HOCl/ClO^- + NO_<span class="hlt">2</span>^-, as reflected in the ratio kII_a(pH 5.5, HOCl)/kII_b(pH 10.0, ClO^-): in H_<span class="hlt">2</span>O (kII_a ˜ 1200 \\cdot kII_b), in 0.1M <span class="hlt">Na</span>Cl (kII_a ˜ 1900 \\cdot kII_b), in 1.0M <span class="hlt">Na</span>Cl (kII_a ˜ 5700 \\cdot kII_b) and in 1.0 M <span class="hlt">Na</span>_<span class="hlt">2</span>SO_4 (kII_a ˜ 1500 \\cdot kII_b). A mechanism for the oxidation of NO</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19970021295&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DTitle%26N%3D0%26No%3D50%26Ntt%3DG','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19970021295&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DTitle%26N%3D0%26No%3D50%26Ntt%3DG"><span>Heterogeneous Reaction of ClONO<span class="hlt">2</span>(g) + <span class="hlt">Na</span>Cl(s) to Cl<span class="hlt">2</span>(g) + <span class="hlt">Na</span>NO3(s)</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Timonen, Raimo S.; Chu, Liang T.; Leu, Ming-Taun; Keyser, Leon F.</p> <p>1994-01-01</p> <p>The heterogeneous reaction of ClON02 + <span class="hlt">Na</span>Cl yields Cl<span class="hlt">2</span> + <span class="hlt">Na</span>NO3 (eq 1) was investigated over a temperature range 220-300 K in a flow-tube reactor interfaced with a differentially pumped quadrupole mass spectrometer. Partial pressures of ClON02 in the range 10(exp -8) - 10(exp -5) Torr were used. Granule sizes and surface roughness of the <span class="hlt">Na</span>Cl substrates were determined by using a scanning electron microscope, and in separate experiments, surface areas of the substrates were measured by using BET analysis of gas-adsorption isotherms. For dry <span class="hlt">Na</span>Cl substrates, both the decay rates of ClON02 and the growth rates Of C12 were used to obtain reaction probabilities, gamma(sub l) = (4.6 +/- 3.0) x 10(exp -3) at 296 K and (6.7 +/- 3.<span class="hlt">2</span>) x 10(exp -1) at 225 K, after considering the internal surface area, The error bars represent 1 standard deviation. The Cl<span class="hlt">2</span> yield based on the ClONO<span class="hlt">2</span> reacted was measured to be 1.0 +/- 0.<span class="hlt">2</span>. In order to mimic the conditions encountered in the lower stratosphere, the effect of water vapor pressures between 5 x 10(exp -5) and 3 x 10(exp -4) Torr on reaction 1 was also studied. With added H20, reaction probabilities, gamma = (4.1 +/- <span class="hlt">2</span>.1) x 10(exp -3) at 296 K and (4.7 +/- <span class="hlt">2</span>.9) x 10(exp -3) at 225 K, were obtained. A trace of HOCl, the reaction product from the ClON02 + H20 yield HOCl + HN03 reaction, was observed in addition to the C12 product from reaction 1. The implications of this result for the enhancement of hydrogen chloride in the stratosphere after the El Chichon volcanic eruption and for the marine troposphere are discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/7026867-sepsis-does-alter-red-blood-cell-glucose-metabolism-na+-concentration-nmr-study','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/7026867-sepsis-does-alter-red-blood-cell-glucose-metabolism-na+-concentration-nmr-study"><span>Sepsis does not alter red blood cell glucose metabolism or <span class="hlt">Na</span>+ concentration: A <span class="hlt">2</span>H-, 23<span class="hlt">Na</span>-NMR study</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>Hotchkiss, R.S.; Song, S.K.; Ling, C.S.</p> <p></p> <p>The effects of sepsis on intracellular <span class="hlt">Na</span>+ concentration ((<span class="hlt">Na</span>+)i) and glucose metabolism were examined in rat red blood cells (RBCs) by using 23<span class="hlt">Na</span>- and <span class="hlt">2</span>H-nuclear magnetic resonance (NMR) spectroscopy. Sepsis was induced in 15 halothane-anesthetized female Sprague-Dawley rats by using the cecal ligation and perforation technique; 14 control rats underwent cecal manipulation without ligation. The animals were fasted for 36 h, but allowed free access to water. At 36 h postsurgery, RBCs were examined by 23<span class="hlt">Na</span>-NMR by using dysprosium tripolyphosphate as a chemical shift reagent. Human RBCs from 17 critically ill nonseptic patients and from 7 patients who were diagnosedmore » as septic were also examined for (<span class="hlt">Na</span>+)i. Five rat RBC specimens had (<span class="hlt">Na</span>+)i determined by both 23<span class="hlt">Na</span>-NMR and inductively coupled plasma-atomic emission spectroscopy (ICP-AES). For glucose metabolism studies, RBCs from septic and control rats were suspended in modified Krebs-Henseleit buffer containing (6,6-<span class="hlt">2</span>H<span class="hlt">2</span>)glucose and examined by <span class="hlt">2</span>H-NMR. No significant differences in (<span class="hlt">Na</span>+)i or glucose utilization were found in RBCs from control or septic rats. There were no differences in (<span class="hlt">Na</span>+)i in the two groups of patients. The (<span class="hlt">Na</span>+)i determined by NMR spectroscopy agreed closely with measurements using ICP-AES and establish that 100% of the (<span class="hlt">Na</span>+)i of the RBC is visible by NMR. Glucose measurements determined by <span class="hlt">2</span>H-NMR correlated closely (correlation coefficient = 0.93) with enzymatic analysis. These studies showed no evidence that sepsis disturbed RBC membrane function or metabolism.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1175751','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1175751"><span>Modulation of contraction by intracellular <span class="hlt">Na</span>+ via <span class="hlt">Na</span>(+)-Ca<span class="hlt">2</span>+ exchange in single shark (Squalus acanthias) ventricular myocytes.</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Näbauer, M; Morad, M</p> <p>1992-01-01</p> <p>1. The effect of direct alteration of intracellular <span class="hlt">Na</span>+ concentration on contractile properties of whole-cell clamped shark ventricular myocytes was studied using an array of 256 photodiodes to monitor the length of the isolated myocytes. <span class="hlt">2</span>. In myocytes dialysed with <span class="hlt">Na</span>(+)-free solution, the voltage dependence of Ca<span class="hlt">2</span>+ current (ICa) and contraction were similar and bell shaped. Contractions activated at all voltages were completely suppressed by nifedipine (5 microM), and failed to show significant tonic components, suggesting dependence of the contraction on Ca<span class="hlt">2</span>+ influx through the L-type Ca<span class="hlt">2</span>+ channel. 3. In myocytes dialysed with 60 mM <span class="hlt">Na</span>+, a ICa-dependent and a ICa-independent component of contraction could be identified. The Ca<span class="hlt">2</span>+ current-dependent component was prominent in voltages between -30 to +10 mV. The ICa-independent contractions were maintained for the duration of depolarization, increased with increasing depolarization between +10 to +100 mV, and were insensitive to nifedipine. 4. In such myocytes, repolarization produced slowly decaying inward tail currents closely related to the time course of relaxation and the degree of shortening prior to repolarization. 5. With 60 mM <span class="hlt">Na</span>+ in the pipette solution, positive clamp potentials activated decaying outward currents which correlated to the size of contraction. These outward currents appeared to be generated by the <span class="hlt">Na</span>(+)-Ca(<span class="hlt">2</span>+)-exchanger since they depended on the presence of intracellular <span class="hlt">Na</span>+, and were neither suppressed by nifedipine nor by K+ channel blockers. 6. The results suggest that in shark (Squalus acanthias) ventricular myocytes, which lack functionally relevant Ca<span class="hlt">2</span>+ release pools, both Ca<span class="hlt">2</span>+ channel and the <span class="hlt">Na</span>(+)-Ca<span class="hlt">2</span>+ exchanger deliver sufficient Ca<span class="hlt">2</span>+ to activate contraction, though the effectiveness of the latter mechanism was highly dependent on the [<span class="hlt">Na</span>+]i. PMID:1338467</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1301138','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1301138"><span><span class="hlt">Na</span>(+) transport, and the E(1)P-E(<span class="hlt">2</span>)P conformational transition of the <span class="hlt">Na</span>(+)/K(+)-ATPase.</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Babes, A; Fendler, K</p> <p>2000-01-01</p> <p>We have used admittance analysis together with the black lipid membrane technique to analyze electrogenic reactions within the <span class="hlt">Na</span>(+) branch of the reaction cycle of the <span class="hlt">Na</span>(+)/K(+)-ATPase. ATP release by flash photolysis of caged ATP induced changes in the admittance of the compound membrane system that are associated with partial reactions of the <span class="hlt">Na</span>(+)/K(+)-ATPase. Frequency spectra and the <span class="hlt">Na</span>(+) dependence of the capacitive signal are consistent with an electrogenic or electroneutral E(1)P <--> E(<span class="hlt">2</span>)P conformational transition which is rate limiting for a faster electrogenic <span class="hlt">Na</span>(+) dissociation reaction. We determine the relaxation rate of the rate-limiting reaction and the equilibrium constants for both reactions at pH 6.<span class="hlt">2</span>-8.5. The relaxation rate has a maximum value at pH 7.4 (approximately 320 s(-1)), which drops to acidic (approximately 190 s(-1)) and basic (approximately 110 s(-1)) pH. The E(1)P <--> E(<span class="hlt">2</span>)P equilibrium is approximately at a midpoint position at pH 6.<span class="hlt">2</span> (equilibrium constant approximately 0.8) but moves more to the E(1)P side at basic pH 8.5 (equilibrium constant approximately 0.4). The <span class="hlt">Na</span>(+) affinity at the extracellular binding site decreases from approximately 900 mM at pH 6.<span class="hlt">2</span> to approximately 200 mM at pH 8.5. The results suggest that during <span class="hlt">Na</span>(+) transport the free energy supplied by the hydrolysis of ATP is mainly used for the generation of a low-affinity extracellular <span class="hlt">Na</span>(+) discharge site. Ionic strength and lyotropic anions both decrease the relaxation rate. However, while ionic strength does not change the position of the conformational equilibrium E(1)P <--> E(<span class="hlt">2</span>)P, lyotropic anions shift it to E(1)P. PMID:11053130</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/2018JPS...382..144B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JPS...382..144B"><span>Sodium intercalation in the phosphosulfate cathode <span class="hlt">Na</span>Fe<span class="hlt">2</span>(PO4)(SO4)<span class="hlt">2</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ben Yahia, Hamdi; Essehli, Rachid; Amin, Ruhul; Boulahya, Khalid; Okumura, Toyoki; Belharouak, Ilias</p> <p>2018-04-01</p> <p>The compound <span class="hlt">Na</span>Fe<span class="hlt">2</span>(PO4)(SO4)<span class="hlt">2</span> is successfully synthesized via a solid state reaction route and its crystal structure is determined using powder X-ray diffraction data. <span class="hlt">Na</span>Fe<span class="hlt">2</span>(PO4)(SO4)<span class="hlt">2</span> phase is also characterized by cyclic voltammetry, galvanostatic cycling and electrochemical impedance spectroscopy. <span class="hlt">Na</span>Fe<span class="hlt">2</span>(PO4)(SO4)<span class="hlt">2</span> crystallizes with the well-known NASICON-type structure. SAED and HRTEM experiments confirm the structural model, and no ordering between the PO4-3 and SO4-<span class="hlt">2</span> polyanions is detected. The electrochemical tests indicate that <span class="hlt">Na</span>Fe<span class="hlt">2</span>(PO4)(SO4)<span class="hlt">2</span> is a 3 V sodium intercalating cathode. The electrical conductivity is relatively low (<span class="hlt">2.2</span> × 10-6 Scm-1 at 200 °C) and the obtained activation energy is ∼0.60eV. The GITT experiments indicate that the diffusivity values are in the range of 10-11-10-12 cm<span class="hlt">2</span>/s within the measured sodium concentrations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017ChJOL..35..681J','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017ChJOL..35..681J"><span>Effect of Ca(OH)<span class="hlt">2</span>, <span class="hlt">Na</span>Cl, and <span class="hlt">Na</span><span class="hlt">2</span>SO4 on the corrosion and electrochemical behavior of rebar</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Jin, Zuquan; Zhao, Xia; Zhao, Tiejun; Hou, Baorong; Liu, Ying</p> <p>2017-05-01</p> <p>The corrosion of rebar in reinforced concrete in marine environments causes significant damage to structures built in ocean environments. Studies on the process and mechanism of corrosion of rebar in the presence of multiple ions may help to control damage and predict the service life of reinforced concrete structures in such environments. The effect of interactions between sulfate and chloride ions and calcium hydroxide on the electrochemical behavior of rebar are also important for evaluation of structure durability. In this work, electrochemical impedance spectroscopy (EIS) plots of rebar in Ca(OH)<span class="hlt">2</span> solution and cement grout, including <span class="hlt">Na</span>Cl and <span class="hlt">Na</span><span class="hlt">2</span>SO4 as aggressive salts, were measured for diff erent immersion times. The results show that corrosion of rebar was controlled by the rate of charge transfer as the rebar was exposed to chloride solution. In the presence of high concentrations of sulfate ions in the electrolyte, generation and dissolution of the passive film proceeded simultaneously and corrosion was mainly controlled by the diff usion rate. When <span class="hlt">Na</span><span class="hlt">2</span>SO4 and <span class="hlt">Na</span>Cl were added to Ca(OH)<span class="hlt">2</span> solution, the instantaneous corrosion rate decreased by a factor of 10 to 20 as a result of the higher pH of the corroding solution.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2230698','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2230698"><span>Role of <span class="hlt">Na</span>+ conductance, <span class="hlt">Na</span>+-H+ exchange, and <span class="hlt">Na</span>+-K+-<span class="hlt">2</span>Cl− symport in the regulatory volume increase of rat hepatocytes</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Wehner, Frank; Tinel, Hanna</p> <p>1998-01-01</p> <p>In rat hepatocytes under hypertonic stress, the entry of <span class="hlt">Na</span>+ (which is thereafter exchanged for K+ via <span class="hlt">Na</span>+-K+-ATPase) plays the key role in regulatory volume increase (RVI).In the present study, the contributions of <span class="hlt">Na</span>+ conductance, <span class="hlt">Na</span>+-H+ exchange and <span class="hlt">Na</span>+-K+-<span class="hlt">2</span>Cl− symport to this process were quantified in confluent primary cultures by means of intracellular microelectrodes and cable analysis, microfluorometric determinations of cell pH and buffer capacity, and measurements of frusemide (furosemide)/bumetanide-sensitive 86Rb+ uptake, respectively. Osmolarity was increased from 300 to 400 mosmol l−1 by addition of sucrose.The experiments indicate a relative contribution of approximately 4:1:1 to hypertonicity-induced <span class="hlt">Na</span>+ entry for the above-mentioned transporters and the overall <span class="hlt">Na</span>+ yield equalled 51 mmol l−1 (10 min)−1.This <span class="hlt">Na</span>+ gain is in good agreement with the stimulation of <span class="hlt">Na</span>+ extrusion via <span class="hlt">Na</span>+-K+-ATPase plus the actual increase in cell <span class="hlt">Na</span>+, namely 55 mmol l−1 (10 min)−1, as was determined on the basis of ouabain-sensitive 86Rb+ uptake and by means of <span class="hlt">Na</span>+-sensitive microelectrodes, respectively.The overall increase in <span class="hlt">Na</span>+ and K+ activity plus the expected concomitant increase in cell Cl− equalled 68 mmol l−1, which fits well with the increase in osmotic activity expected to occur from an initial cell shrinkage to 87.5 % and a RVI to 92.6 % of control, namely 53 mosmol l−1.The prominent role of <span class="hlt">Na</span>+ conductance in the RVI of rat hepatocytes could be confirmed on the basis of the pharmacological profile of this process, which was characterized by means of confocal laser-scanning microscopy. PMID:9481677</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28391722','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28391722"><span>Eversion Strength and Surface Electromyography Measures With and Without Chronic Ankle Instability Measured in <span class="hlt">2</span> Positions.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Donnelly, Lindsy; Donovan, Luke; Hart, Joseph M; Hertel, Jay</p> <p>2017-07-01</p> <p>Individuals with chronic ankle instability (<span class="hlt">CAI</span>) have demonstrated strength deficits compared to healthy controls; however, the influence of ankle position on force measures and surface electromyography (sEMG) activation of the peroneus longus and brevis has not been investigated. The purpose of this study was to compare sEMG amplitudes of the peroneus longus and brevis and eversion force measures in <span class="hlt">2</span> testing positions, neutral and plantarflexion, in groups with and without <span class="hlt">CAI</span>. Twenty-eight adults (19 females, 9 males) with <span class="hlt">CAI</span> and 28 healthy controls (19 females, 9 males) participated. Hand-held dynamometer force measures were assessed during isometric eversion contractions in <span class="hlt">2</span> testing positions (neutral, plantarflexion) while surface sEMG amplitudes of the peroneal muscles were recorded. Force measures were normalized to body mass, and sEMG amplitudes were normalized to a resting period. The group with <span class="hlt">CAI</span> demonstrated less force when compared to the control group ( P < .001) in both the neutral and plantarflexion positions: neutral position, <span class="hlt">CAI</span>: 1.64 Nm/kg and control: <span class="hlt">2</span>.10 Nm/kg) and plantarflexion position, <span class="hlt">CAI</span>: 1.40 Nm/kg and control: 1.73 Nm/kg). There were no differences in sEMG amplitudes between the groups or muscles ( P > .05). Force measures correlated with both muscles' sEMG amplitudes in the healthy group (neutral peroneus longus: r = 0.42, P = .03; plantarflexion peroneus longus: r = 0.56, P = .002; neutral peroneus brevis: r = 0.38, P = .05; plantarflexion peroneus longus: r = 0.40, P = .04), but not in the group with <span class="hlt">CAI</span> ( P > .05). The group with <span class="hlt">CAI</span> generated less force when compared to the control group during both testing positions. There was no selective activation of the peroneal muscles with testing in both positions, and force output and sEMG activity was only related in the healthy group. Clinicians should assess eversion strength and implement strength training exercises in different sagittal plane positions and evaluate for other</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016ApSS..378..167Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016ApSS..378..167Z"><span>Combined effects <span class="hlt">Na</span> and SO<span class="hlt">2</span> in flue gas on Mn-Ce/TiO<span class="hlt">2</span> catalyst for low temperature selective catalytic reduction of NO by NH3 simulated by <span class="hlt">Na</span><span class="hlt">2</span>SO4 doping</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zhou, Aiyi; Yu, Danqing; Yang, Liu; Sheng, Zhongyi</p> <p>2016-08-01</p> <p>A series of Mn-Ce/TiO<span class="hlt">2</span> catalysts were synthesized through an impregnation method and used for low temperature selective catalytic reduction (SCR) of NOx with ammonia (NH3). <span class="hlt">Na</span><span class="hlt">2</span>SO4 was added into the catalyst to simulate the combined effects of alkali metal and SO<span class="hlt">2</span> in the flue gas. Experimental results showed that <span class="hlt">Na</span><span class="hlt">2</span>SO4 had strong and fluctuant influence on the activity of Mn-Ce/TiO<span class="hlt">2</span>, because the effect of <span class="hlt">Na</span><span class="hlt">2</span>SO4 included pore occlusion and sulfation effect simultaneously. When <span class="hlt">Na</span><span class="hlt">2</span>SO4 loading content increased from 0 to 1 wt.%, the SCR activities of <span class="hlt">Na</span><span class="hlt">2</span>SO4-doped catalysts decreased greatly. With further increasing amount of <span class="hlt">Na</span><span class="hlt">2</span>SO4, however, the catalytic activity increased gradually. XRD results showed that <span class="hlt">Na</span><span class="hlt">2</span>SO4 doping could induce the crystallization of MnOx phases, which were also confirmed by TEM and SEM results. BET results showed that the surface areas decreased and a new bimodal mesoporous structure formed gradually with the increasing amount of <span class="hlt">Na</span><span class="hlt">2</span>SO4. XPS results indicated that part of Ce4+ and Mn3+ were transferred to Ce3+ and Mn4+ due to the sulfation after <span class="hlt">Na</span><span class="hlt">2</span>SO4 deposition on the surface of the catalysts. When the doped amounts of <span class="hlt">Na</span><span class="hlt">2</span>SO4 increased, NH3-TPD results showed that the Lewis acid sites decreased and the Brønsted acid sites of Mn-Ce/TiO<span class="hlt">2</span> increased quickly, which could be considered as another reason for the observed changes in the catalytic activity. The decreased Mn and Ce atomic concentration, the changes of their oxidative states, and the variation in acidic properties on the surface of <span class="hlt">Na</span><span class="hlt">2</span>SO4-doped catalysts could be the reasons for the fluctuant changes of the catalytic activity.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25686329','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25686329"><span>[Synergistic effects of lysozyme with EDTA-<span class="hlt">2</span><span class="hlt">Na</span> on antibacterial activity].</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Li, Xiao-man; Wang, Xiao-yan; Gao, Xue-jun</p> <p>2015-02-18</p> <p>To evaluate the synergistic antibacterial effects of lysozyme with ethylenediaminetetraacetic acid disodium salt (EDTA-<span class="hlt">2</span><span class="hlt">Na</span>) on Enterococcus faecalis (E. faecalis) and Porphyromonas endodontalis (P. endodontalis). E. faecalis and P. endodontalis were cultured and adjusted to 10(8) CFU/mL. Then 0.3, 0.5, 1, <span class="hlt">2</span>, 5, 10, 50, 100, 150 and 300 g/L of lysozyme were prepared with deionized water; and the lysozyme solutions were mixed with 0.5, 1.0, <span class="hlt">2</span>.0 g/L of EDTA-<span class="hlt">2</span><span class="hlt">Na</span>, respectively. The bacteria and lysosome with/without EDTA-<span class="hlt">2</span><span class="hlt">Na</span> interacted for 15 min, then water-soluble tetrazolium (WST) working solution was added and the activity of the bacteria was calculated by measuring optical densities at 450 nm and 630 nm with microplate spectrophotometer. Regarding the pure lysozyme from 0.5 g/L to 150 g/L, more E. faecalis and P. endodontalis were inhibited when the concentration of lysozyme was higher, especially for E. faecalis. There was synergistic effect of lysozyme with EDTA-<span class="hlt">2</span><span class="hlt">Na</span> on antibacterial activity, which was related to the concentration of lysozyme. On E. faecalis, the antibacterial activity of lysozyme with EDTA-<span class="hlt">2</span><span class="hlt">Na</span> was 1.<span class="hlt">2</span>-3.7 folds than the pure lysozyme when the concentration of lysozyme was 0.5-50 g/L (P<0.05), and on P. endodontalis, the antibacterial activity of lysozyme with EDTA-<span class="hlt">2</span><span class="hlt">Na</span> was 1.3-3.5 folds than the pure lysozyme when the concentration of lysozyme was 0.5-10 g/L (P<0.05). When the concentration of lysozyme was higher than 100 g/L, EDTA-<span class="hlt">2</span><span class="hlt">Na</span> did not show synergistic effect on the antibacterial activity (P>0.05). For E. faecalis and P. endodontalis, a low concentration of lysozyme with EDTA-<span class="hlt">2</span><span class="hlt">Na</span> showed significant synergistic antibacterial activity, while a high concentration of lysozyme with EDTA-<span class="hlt">2</span><span class="hlt">Na</span> did not.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AIPC.1788c0066B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AIPC.1788c0066B"><span>Hot-corrosion of AISI 1020 steel in a molten <span class="hlt">NaCl/Na</span><span class="hlt">2</span>SO4 eutectic at 700°C</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Badaruddin, Mohammad; Risano, Ahmad Yudi Eka; Wardono, Herry; Asmi, Dwi</p> <p>2017-01-01</p> <p>Hot-corrosion behavior and morphological development of AISI 1020 steel with <span class="hlt">2</span> mg cm-<span class="hlt">2</span> mixtures of various <span class="hlt">NaCl/Na</span><span class="hlt">2</span>SO4 ratios at 700°C were investigated by means of weight gain measurements, Optical Microscope (OM), X-ray diffraction (XRD), scanning electron microscopy (SEM), and energy dispersive X-ray spectroscopy (EDS). The weight gain kinetics of the steel with mixtures of salt deposits display a rapid growth rates, compared with the weight gain kinetics of AISI 1020 steel without salt deposit in dry air oxidation, and follow a steady-state parabolic law for 49 h. Chloridation and sulfidation produced by a molten <span class="hlt">NaCl/Na</span><span class="hlt">2</span>SO4 on the steel induced hot-corrosion mechanism attack, and are responsible for the formation of thicker scale. The most severe corrosion takes place with the 70 wt.% <span class="hlt">Na</span>Cl mixtures in <span class="hlt">Na</span><span class="hlt">2</span>SO4. The typical Fe<span class="hlt">2</span>O3 whisker growth in outer part scale was attributed to the FeCl3 volatilization. The formation of FeS in the innermost scale is more pronounced as the content of <span class="hlt">Na</span><span class="hlt">2</span>SO4 in the mixture is increased.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70019363','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70019363"><span>Fluor-ferro-leakeite, <span class="hlt">NaNa</span><span class="hlt">2</span>(FC<span class="hlt">2+2</span>Fe3+<span class="hlt">2</span>Li)Si8O22F<span class="hlt">2</span>, a new alkali amphibole from the Canada Pinabete pluton, Questa, New Mexico, U.S.A.</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Hawthorne, F.C.; Oberti, R.; Ungaretti, L.; Ottolini, L.; Grice, Joel D.; Czamanske, G.K.</p> <p>1996-01-01</p> <p>Fluor-ferro-leakeite is a new amphibole species from the Canada Pinabete pluton, Questa, New Mexico, U.S.A.; it occurs in association with quartz, alkali feldspar, acmite, ilmenite, and zircon. It forms as anhedral bluish black crystals elongated along c and up to 1 mm long. It is brittle, H = 6, Dmeas = 3.37 g/cm3, Dcalc = 3.34 g/cm3. In plane-polarized light, it is strongly pleochroic, X = very dark indigo blue, Y = gray blue, Z = yellow green; X ??? c = 10?? (in ??obtuse), Y = b, Z ??? a = 4?? (in ?? obtuse), with absorption X > Y > Z. Fluor-ferro-leakeite is biaxial positive, ?? = 1.675(<span class="hlt">2</span>), ??= 1.683(<span class="hlt">2</span>), ?? = 1.694(1); <span class="hlt">2</span>V = 87(<span class="hlt">2</span>)??; dispersion is not visible because of the strong absorption. Fluor-ferro-leakeite is monoclinic, space group C<span class="hlt">2</span>/m, a = 9.792(1), b = 17.938(1), c = 5.3133(4) A??, ??= 103.87(7)??, V = 906.0(1) A??3, Z = <span class="hlt">2</span>. The ten strongest X-ray diffraction lines in the powder pattern are [d(I,hkl)]: <span class="hlt">2</span>.710(100,151), <span class="hlt">2</span>.536(92,202), 3.404(57,131), 4.481(54,040), 8.426(45,110), <span class="hlt">2</span>.985(38,241), <span class="hlt">2</span>.585(38,061), 3.122(29,310), <span class="hlt">2</span>.165(26,261), and 1.586(25,403). Analysis by a combination of electron microprobe, ion microprobe, and crystal-structure refinement (Hawthorne et al. 1993) gives SiO<span class="hlt">2</span> 51.12, Al<span class="hlt">2</span>O3 1.13, TiO<span class="hlt">2</span> 0.68, Fe<span class="hlt">2</span>O3 16.73, FeO 8.87, MgO <span class="hlt">2</span>.02, MnO 4.51, ZnO 0.57, CaO 0.15, <span class="hlt">Na</span><span class="hlt">2</span>O 9.22, K<span class="hlt">2</span>O 1.19, Li<span class="hlt">2</span>O 0.99, F <span class="hlt">2</span>.87, H<span class="hlt">2</span>Ocalc 0.60, sum 99.44 wt%. The formula unit, calculated on the basis of 23 O atoms, is (K0.23<span class="hlt">Na</span>0.76)(<span class="hlt">Na</span>1.97Ca0.03)(Mg 0.46Fe<span class="hlt">2</span>+1.4Mn<span class="hlt">2</span>+0.59Zn0.07Fe3+1.93-Ti 0.08Al0.02Li0.61])(Si7.81Al 0.19)O22(F1.39OH0.61). A previous crystal-structure refinement (Hawthorne et al. 1993) shows Li to be completely ordered at the M3 site. Fluor-ferro-leakeite, ideally <span class="hlt">NaNa</span><span class="hlt">2</span>(Fe<span class="hlt">2+2</span>Fe3+<span class="hlt">2</span>Li)Si8O22F<span class="hlt">2</span>, is related to leakeite, <span class="hlt">NaNa</span><span class="hlt">2</span>(Mg<span class="hlt">2</span>Fe3+3Li)Si 8O22(OH)<span class="hlt">2</span>, by the substitutions Fe<span class="hlt">2</span>+ ??? Mg and F ??? OH.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29497735','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29497735"><span>Thermodynamic description of Tc(iv) solubility and carbonate complexation in alkaline <span class="hlt">Na</span>HCO3-<span class="hlt">Na</span><span class="hlt">2</span>CO3-<span class="hlt">Na</span>Cl systems.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Baumann, A; Yalçıntaş, E; Gaona, X; Polly, R; Dardenne, K; Prüßmann, T; Rothe, J; Altmaier, M; Geckeis, H</p> <p>2018-03-28</p> <p>The solubility of 99 Tc(iv) was investigated in dilute to concentrated carbonate solutions (0.01 M ≤ C tot ≤ 1.0 M, with C tot = [HCO 3 - ] + [CO 3 <span class="hlt">2</span>- ]) under systematic variation of ionic strength (I = 0.3-5.0 M <span class="hlt">Na</span>HCO 3 -<span class="hlt">Na</span> <span class="hlt">2</span> CO 3 -<span class="hlt">NaCl-Na</span>OH) and pH m (-log[H + ] = 8.5-14.5). Strongly reducing conditions (pe + pH m ≈ <span class="hlt">2</span>) were set with Sn(ii). Carbonate enhances the solubility of Tc(iv) in alkaline conditions by up to 3.5 log 10 -units compared to carbonate-free systems. Solvent extraction and XANES confirmed that Tc was kept as +IV during the timeframe of the experiments (≤ 650 days). Solid phase characterization performed by XAFS, XRD, SEM-EDS, chemical analysis and TG-DTA confirmed that TcO <span class="hlt">2</span> ·0.6H <span class="hlt">2</span> O(am) controls the solubility of Tc(iv) under the conditions investigated. Slope analysis of the solubility data in combination with solid/aqueous phase characterization and DFT calculations indicate the predominance of the species Tc(OH) 3 CO 3 - at pH m ≤ 11 and C tot ≥ 0.01 M, for which thermodynamic and activity models are derived. Solubility data obtained above pH m ≈ 11 indicates the formation of previously unreported Tc(iv)-carbonate species, possibly Tc(OH) 4 CO 3 <span class="hlt">2</span>- , although the likely formation of additional complexes prevents deriving a thermodynamic model valid for this pH m -region. This work provides the most comprehensive thermodynamic dataset available for the system Tc 4+ -<span class="hlt">Na</span> + -Cl - -OH - -HCO 3 - -CO 3 <span class="hlt">2</span>- -H <span class="hlt">2</span> O(l) valid under a range of conditions relevant for nuclear waste disposal.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018EML....14...30S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018EML....14...30S"><span>Fabrication of <span class="hlt">Na</span>0.7MnO<span class="hlt">2</span>/C composite cathode material by simple heat treatment for high-power <span class="hlt">na</span>-ion batteries</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sohn, DongRak; Lim, Sung-Jin; Nam, Do-Hwan; Hong, Kyung-Sik; Kim, Tae-Hee; Oh, SeKwon; Eom, Ji-Yong; Cho, EunAe; Kwon, HyukSang</p> <p>2018-01-01</p> <p>A <span class="hlt">Na</span>0.7MnO<span class="hlt">2</span>/C composite cathode material is synthesized by simple and costeffective two-step heat treatment for an improvement in the rate capability of <span class="hlt">Na</span>0.7MnO<span class="hlt">2</span>. The first heat treatment is to synthesize <span class="hlt">Na</span>0.7MnO<span class="hlt">2</span>, and the second one is a low temperature annealing at 350 °C for 1 h in air, which is necessary to suppress an interfacial reaction between the <span class="hlt">Na</span>0.7MnO<span class="hlt">2</span> and C in the synthesis process of <span class="hlt">Na</span>0.7MnO<span class="hlt">2</span>/C composite. Structural analyses by XRD and XPS reveal that the <span class="hlt">Na</span>0.7MnO<span class="hlt">2</span>/C shows the same structural properties as that of the pristine <span class="hlt">Na</span>0.7MnO<span class="hlt">2</span>, and hence they exhibit the same initial discharge capacity of 175 mAh g-1 at 20 mA g-1. At a current density of 400 mA g-1, the discharge capacity of <span class="hlt">Na</span>0.7MnO<span class="hlt">2</span> reduces to 50 mAh g-1 (28% of the initial discharge capacity), whereas that of <span class="hlt">Na</span>0.7MnO<span class="hlt">2</span>/C reduces to 108 mAh g-1 (61% of the initial discharge capacity). The enhanced rate capability of the <span class="hlt">Na</span>0.7MnO<span class="hlt">2</span>/C is attributed to the conductive carbon layer formed on the surface of <span class="hlt">Na</span>0.7MnO<span class="hlt">2</span> particles, enabling the facile transport of electrons from the current collector to the surface of the <span class="hlt">Na</span>0.7MnO<span class="hlt">2</span> particles. [Figure not available: see fulltext.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1332788','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1332788"><span>Robust high pressure stability and negative thermal expansion in sodium-rich antiperovskites <span class="hlt">Na</span> 3OBr and <span class="hlt">Na</span> 4OI <span class="hlt">2</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>Wang, Yonggang; Wen, Ting; Park, Changyong</p> <p>2016-01-14</p> <p>The structure stability under high pressure and thermal expansion behavior of <span class="hlt">Na</span> 3OBr and <span class="hlt">Na</span> 4OI <span class="hlt">2</span>, two prototypes of alkali-metal-rich antiperovskites, were investigated by in situ synchrotron X-ray diffraction techniques under high pressure and low temp. Both are soft materials with bulk modulus of 58.6 GPa and 52.0 GPa for <span class="hlt">Na</span> 3OBr and <span class="hlt">Na</span> 4OI <span class="hlt">2</span>, resp. The cubic <span class="hlt">Na</span> 3OBr structure and tetragonal <span class="hlt">Na</span> 4OI <span class="hlt">2</span> with intergrowth K <span class="hlt">2</span>NiF 4 structure are stable under high pressure up to 23 GPa. Although being a characteristic layered structure, <span class="hlt">Na</span> 4OI <span class="hlt">2</span> exhibits nearly isotropic compressibility. Neg. thermal expansion wasmore » obsd. at low temp. range (20-80 K) in both transition-metal-free antiperovskites for the first time. The robust high pressure structure stability was examined. and confirmed by first-principles calculations. among various possible polymorphisms qualitatively. The results provide in-depth understanding of the neg. thermal expansion and robust crystal structure stability of these antiperovskite systems and their potential applications.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/1393849-structure-electrochemical-evolution-mn-rich-p2-na2-na-ion-battery-cathode','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/1393849-structure-electrochemical-evolution-mn-rich-p2-na2-na-ion-battery-cathode"><span>Structure-electrochemical evolution of a Mn-rich P<span class="hlt">2</span> <span class="hlt">Na</span> <span class="hlt">2</span>/3Fe 0.<span class="hlt">2</span>Mn 0.8O <span class="hlt">2</span> <span class="hlt">Na</span>-ion battery cathode</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>Dose, Wesley M.; Sharma, Neeraj; Pramudita, James C.</p> <p></p> <p>The structural evolution of electrode materials directly influences the performance of sodium-ion batteries. In this work, in situ synchrotron X-ray diffraction is used to investigate the evolution of the crystal structure of a Mn-rich P<span class="hlt">2</span>-phase <span class="hlt">Na</span> <span class="hlt">2</span>/3Fe 0.<span class="hlt">2</span>Mn 0.8O <span class="hlt">2</span> cathode. A single-phase reaction takes place for the majority of the discharge–charge cycle at ~C/10, with only a short, subtle hexagonal P<span class="hlt">2</span> to hexagonal P<span class="hlt">2</span> two-phase region early in the first charge. Thus, a higher fraction of Mn compared to previous studies is demonstrated to stabilize the P<span class="hlt">2</span> structure at high and low potentials, with neither “Z”/OP4 phases in themore » charged state nor significant quantities of the P'<span class="hlt">2</span> phase in the discharged state between 1.5 and 4.<span class="hlt">2</span> V. Notably, sodium ions inserted during discharge are located on both available crystallographic sites, albeit with a preference for the site sharing edges with the MO 6 octahedral unit. The composition <span class="hlt">Na</span> ~0.70Fe 0.<span class="hlt">2</span>Mn 0.8O <span class="hlt">2</span> prompts a reversible single-phase sodium redistribution between the two sites. Sodium ions vacate the site sharing faces (Naf), favoring the site sharing edges (Nae) to give a Nae/Naf site occupation of 4:1 in the discharged state. This site preference could be an intermediate state prior to the formation of the P'<span class="hlt">2</span> phase. Furthermore, this work shows how the Mn-rich <span class="hlt">Na</span> <span class="hlt">2</span>/3Fe 0.<span class="hlt">2</span>Mn 0.8O <span class="hlt">2</span> composition and its sodium-ion distribution can minimize phase transitions during battery function, especially in the discharged state.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1393849-structure-electrochemical-evolution-mn-rich-p2-na2-na-ion-battery-cathode','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1393849-structure-electrochemical-evolution-mn-rich-p2-na2-na-ion-battery-cathode"><span>Structure-electrochemical evolution of a Mn-rich P<span class="hlt">2</span> <span class="hlt">Na</span> <span class="hlt">2</span>/3Fe 0.<span class="hlt">2</span>Mn 0.8O <span class="hlt">2</span> <span class="hlt">Na</span>-ion battery cathode</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Dose, Wesley M.; Sharma, Neeraj; Pramudita, James C.; ...</p> <p>2017-08-04</p> <p>The structural evolution of electrode materials directly influences the performance of sodium-ion batteries. In this work, in situ synchrotron X-ray diffraction is used to investigate the evolution of the crystal structure of a Mn-rich P<span class="hlt">2</span>-phase <span class="hlt">Na</span> <span class="hlt">2</span>/3Fe 0.<span class="hlt">2</span>Mn 0.8O <span class="hlt">2</span> cathode. A single-phase reaction takes place for the majority of the discharge–charge cycle at ~C/10, with only a short, subtle hexagonal P<span class="hlt">2</span> to hexagonal P<span class="hlt">2</span> two-phase region early in the first charge. Thus, a higher fraction of Mn compared to previous studies is demonstrated to stabilize the P<span class="hlt">2</span> structure at high and low potentials, with neither “Z”/OP4 phases in themore » charged state nor significant quantities of the P'<span class="hlt">2</span> phase in the discharged state between 1.5 and 4.<span class="hlt">2</span> V. Notably, sodium ions inserted during discharge are located on both available crystallographic sites, albeit with a preference for the site sharing edges with the MO 6 octahedral unit. The composition <span class="hlt">Na</span> ~0.70Fe 0.<span class="hlt">2</span>Mn 0.8O <span class="hlt">2</span> prompts a reversible single-phase sodium redistribution between the two sites. Sodium ions vacate the site sharing faces (Naf), favoring the site sharing edges (Nae) to give a Nae/Naf site occupation of 4:1 in the discharged state. This site preference could be an intermediate state prior to the formation of the P'<span class="hlt">2</span> phase. Furthermore, this work shows how the Mn-rich <span class="hlt">Na</span> <span class="hlt">2</span>/3Fe 0.<span class="hlt">2</span>Mn 0.8O <span class="hlt">2</span> composition and its sodium-ion distribution can minimize phase transitions during battery function, especially in the discharged state.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26783178','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26783178"><span>Discovery of <span class="hlt">2</span>-azetidinone and 1H-pyrrole-<span class="hlt">2</span>,5-dione derivatives containing sulfonamide group at the side chain as potential cholesterol absorption inhibitors.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Yuan, Xinrui; Lu, Peng; Xue, Xiaojian; Qin, Hui; Fan, Chen; Wang, Yubin; Zhang, Qi</p> <p>2016-02-01</p> <p>Cholesterol absorption inhibitor (<span class="hlt">CAI</span>) targeting Niemann-Pick C1-like1 protein was developed for the treatment of hyperlipidaemia and only ezetimibe was approved so far. For developing novel <span class="hlt">CAIs</span>, we synthesized sixteen <span class="hlt">2</span>-azetidinone derivatives and thirteen 1H-pyrrole-<span class="hlt">2</span>,5-dione derivatives containing sulfonamide group at the side chain, and their inhibitory activity of cholesterol absorption was evaluated in Caco-<span class="hlt">2</span> cell line in vitro. Furthermore, top six compounds were measured by cytotoxicity and partition coefficient, and <span class="hlt">2</span>-azetidinone analogue 9e was selected for in vivo study. Finally, 9e considerably reduced total cholesterol, LDL-C, FFA and triglyceride in the serum and increased the rate of HDL-C to total cholesterol, suggesting it could regulate the lipid metabolism and act as a potent <span class="hlt">CAI</span>. Copyright © 2015 Elsevier Ltd. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/1157064-thermodynamic-kinetic-analyses-co2-chemisorption-mechanism-na2tio3-experimental-theoretical-evidences','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/1157064-thermodynamic-kinetic-analyses-co2-chemisorption-mechanism-na2tio3-experimental-theoretical-evidences"><span>Thermodynamic and kinetic analyses of the CO<span class="hlt">2</span> chemisorption mechanism on <span class="hlt">Na</span><span class="hlt">2</span>TiO3: Experimental and theoretical evidences</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>Duan, Yuhua</p> <p>2014-01-01</p> <p>ABSTRACT: Sodium metatitanate (<span class="hlt">Na</span><span class="hlt">2</span>TiO3) was successfully synthesized via a solid-state reaction. The <span class="hlt">Na</span><span class="hlt">2</span>TiO3 structure and microstructure were characterized using X-ray diffraction, scanning and transmission electron microscopy, and N<span class="hlt">2</span> adsorption. Then, the CO<span class="hlt">2</span> chemisorption mechanism on <span class="hlt">Na</span><span class="hlt">2</span>TiO3 was systematically analyzed to determine the influence of temperature. The CO<span class="hlt">2</span> chemisorption capacity of <span class="hlt">Na</span><span class="hlt">2</span>TiO3 was evaluated both dynamically and isothermally, and the products were reanalyzed to elucidate the <span class="hlt">Na</span><span class="hlt">2</span>TiO3-CO<span class="hlt">2</span> reaction mechanism. Different chemical species (<span class="hlt">Na</span><span class="hlt">2</span>CO3, <span class="hlt">Na</span><span class="hlt">2</span>O, and <span class="hlt">Na</span>4Ti5O12 or <span class="hlt">Na</span>16Ti10O28) were identified during the CO<span class="hlt">2</span> capture process in <span class="hlt">Na</span><span class="hlt">2</span>TiO3. In addition, some CO<span class="hlt">2</span> chemisorption kinetic parameters were determined. The ΔH‡ was found tomore » be 140.9 kJ/mol, to the <span class="hlt">Na</span><span class="hlt">2</span>TiO3-CO<span class="hlt">2</span> system, between 600 and 780 °C. Results evidenced that CO<span class="hlt">2</span> chemisorption on <span class="hlt">Na</span><span class="hlt">2</span>TiO3 highly depends on the reaction temperature. Furthermore, the experiments were theoretically supported by different thermodynamic calculations. The calculated thermodynamic properties of CO<span class="hlt">2</span> capture reactions by (<span class="hlt">Na</span><span class="hlt">2</span>TiO3, <span class="hlt">Na</span>4Ti5O12, and <span class="hlt">Na</span>16Ti10O28) sodium titanates were fully investigated.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1185927-ft-ir-study-co2-interaction-na-rich-montmorillonite','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1185927-ft-ir-study-co2-interaction-na-rich-montmorillonite"><span>FT-IR study of CO <span class="hlt">2</span> interaction with <span class="hlt">Na</span>-rich montmorillonite</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Krukowski, Elizabeth G.; Goodman, Angela; Rother, Gernot; ...</p> <p>2015-05-27</p> <p>Here, carbon capture, utilization and storage (CCUS) in saline reservoirs in sedimentary formations has the potential to reduce the impact of fossil fuel combustion on climate change by reducing CO <span class="hlt">2</span> emissions to the atmosphere and storing the CO <span class="hlt">2</span> in geologic formations in perpetuity. At pressure and temperature (PT) conditions relevant to CCUS, CO <span class="hlt">2</span> is less dense than the pre-existing brine in the formation, and the more buoyant CO <span class="hlt">2</span> will migrate to the top of the formation where it will be in contact with cap rock. Interactions between clay-rich shale cap rocks and CO <span class="hlt">2</span> are poorlymore » understood at PT conditions appropriate for CCUS in saline formations. In this study, the interaction of CO <span class="hlt">2</span> with clay minerals in the cap rock overlying a saline formation has been examined using <span class="hlt">Na</span> + exchanged montmorillonite (Mt) (<span class="hlt">Na</span> +-STx-1) (<span class="hlt">Na</span> + Mt) as an analog for clay-rich shale. Attenuated Total Reflectance-Fourier Transform Infrared Spectroscopy (ATR-FTIR) was used to discern mechanistic information for CO <span class="hlt">2</span> interaction with hydrated (both one- and two-water layers) and relatively dehydrated (both dehydrated layers and one-water layers) <span class="hlt">Na</span>+-STx-1 at 35 °C and 50 C and CO <span class="hlt">2</span> pressure from 0 5.9 MPa. CO <span class="hlt">2</span>-induced perturbations associated with the water layer and <span class="hlt">Na</span>+-STx-1 vibrational modes such as AlAlOH and AlMgOH were examined. Data indicate that CO <span class="hlt">2</span> is preferentially incorporated into the interlayer space, with relatively dehydrated <span class="hlt">Na</span> +-STx-1 capable of incorporating more CO <span class="hlt">2</span> compared to hydrated <span class="hlt">Na</span> +-STx-1. Spectroscopic data provide no evidence of formation of carbonate minerals or the interaction of CO <span class="hlt">2</span> with sodium cations in the <span class="hlt">Na</span> +-STx-1 structure.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/17680266','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/17680266"><span>Osmolality- and <span class="hlt">Na</span>+ -dependent effects of hyperosmotic <span class="hlt">Na</span>Cl solution on contractile activity and Ca<span class="hlt">2</span>+ cycling in rat ventricular myocytes.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Ricardo, Rafael A; Bassani, Rosana A; Bassani, José W M</p> <p>2008-01-01</p> <p>Hypertonic <span class="hlt">Na</span>Cl solutions have been used for small-volume resuscitation from hypovolemic shock. We sought to identify osmolality- and <span class="hlt">Na</span>(+)-dependent components of the effects of the hyperosmotic <span class="hlt">Na</span>Cl solution (85 mOsm/kg increment) on contraction and cytosolic Ca(<span class="hlt">2</span>+) concentration ([Ca(<span class="hlt">2</span>+)](i)) in isolated rat ventricular myocytes. The biphasic change in contraction and Ca(<span class="hlt">2</span>+) transient amplitude (decrease followed by recovery) was accompanied by qualitatively similar changes in sarcoplasmic reticulum (SR) Ca(<span class="hlt">2</span>+) content and fractional release and was mimicked by isosmotic, equimolar increase in extracellular [<span class="hlt">Na</span>(+)] ([<span class="hlt">Na</span>(+)](o)). Raising osmolality with sucrose, however, augmented systolic [Ca(<span class="hlt">2</span>+)](i) monotonically without change in SR parameters and markedly decreased contraction amplitude and diastolic cell length. Functional SR inhibition with thapsigargin abolished hyperosmolality effects on [Ca(<span class="hlt">2</span>+)](i). After 15-min perfusion, both hyperosmotic solutions slowed mechanical relaxation during twitches and [Ca(<span class="hlt">2</span>+)](i) decline during caffeine-evoked transients, raised diastolic and systolic [Ca(<span class="hlt">2</span>+)](i), and depressed systolic contractile activity. These effects were greater with sucrose solution, and were not observed after isosmotic [<span class="hlt">Na</span>(+)](o) increase. We conclude that under the present experimental conditions, transmembrane <span class="hlt">Na</span>(+) redistribution apparently plays an important role in determining changes in SR Ca(<span class="hlt">2</span>+) mobilization, which markedly affect contractile response to hyperosmotic <span class="hlt">Na</span>Cl solutions and attenuate the osmotically induced depression of contractile activity.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23288200','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23288200"><span>Bergamot essential oil differentially modulates intracellular Ca<span class="hlt">2</span>+ levels in vascular endothelial and smooth muscle cells: a new finding seen with fura-<span class="hlt">2</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>You, Ji H; Kang, Purum; Min, Sun Seek; Seol, Geun Hee</p> <p>2013-04-01</p> <p>In this study, we compared the effect of the essential oil of Citrus bergamia Risso [bergamot, bergamot essential oil (BEO)] on the intracellular Ca levels in vascular endothelial (EA) and mouse vascular smooth muscle (MOVAS) cells, using the fura-<span class="hlt">2</span> fluorescence technique. BEO caused an initial transient increase in intracellular Ca concentration ([<span class="hlt">Ca]i</span>) in EA cells, followed by a decrease, whereas it induced a sustained increase in [<span class="hlt">Ca]i</span> in MOVAS cells. Linalyl acetate (LA) as a major component of BEO-induced [<span class="hlt">Ca]i</span> mobilization was similar to BEO in EA cells. The increase of [<span class="hlt">Ca]i</span> by LA was higher in EA cells than in MOVAS cells. [<span class="hlt">Ca]i</span> rise induced by extracellular Ca application was significantly blocked by BEO or LA in EA cells but not in MOVAS cells, suggesting that BEO and LA block Ca influx in EA cells. The present results suggest that BEO and LA differentially modulate intracellular Ca levels in vascular endothelial and smooth muscle cells. In addition, blockade of Ca influx by BEO and LA in EA cells may explain the protective effects of BEO on endothelial dysfunction associated with cardiovascular disease.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1311298-ph-regulative-synthesis-na3-vpo4-nanoflowers-improved-na-cycling-stability','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1311298-ph-regulative-synthesis-na3-vpo4-nanoflowers-improved-na-cycling-stability"><span>pH-regulative synthesis of <span class="hlt">Na</span> 3(VPO 4) <span class="hlt">2</span>F 3 nanoflowers and their improved <span class="hlt">Na</span> cycling stability</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Qi, Yuruo; Mu, Linqin; Zhao, Junmei; ...</p> <p>2016-04-08</p> <p><span class="hlt">Na</span>-ion batteries are becoming increasingly attractive as a low cost energy storage device. Sodium vanadium fluorophosphates have been studied extensively recently due to their high storage capacity and high discharge voltage. Shape and size often have a crucial influence over the properties. The controlling synthesis of nanoparticles with special microstructures is significant, which becomes a challenging issue and has drawn considerable attention. In this study, <span class="hlt">Na</span> 3(VPO 4) <span class="hlt">2</span>F 3 nanoflowers have been synthesized via a pH-regulative low-temperature (120 °C) hydro-thermal route. In particular, it is a green route without any organic compounds involved. The hydro-thermal reaction time for themore » formation of <span class="hlt">Na</span> 3(VPO 4) <span class="hlt">2</span>F 3 nanoflowers has also been investigated. A weak acid environment (pH = <span class="hlt">2</span>.60) with the possible presence of hydrogen fluoride molecules is necessary for the formation of the desired nanoflower microstructures. Moreover, compared to the nanoparticles obtained by <span class="hlt">Na</span> <span class="hlt">2</span>HPO 4·12H <span class="hlt">2</span>O, the as-synthesized <span class="hlt">Na</span> 3(VPO 4) <span class="hlt">2</span>F 3 nanoflowers showed an excellent <span class="hlt">Na</span>-storage performance in terms of superior cycle stability, even without any further carbon coating or high-temperature treatment.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/1406910-layered-p2-na-co-ti-high-performance-cathode-material-sodium-ion-batteries','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/1406910-layered-p2-na-co-ti-high-performance-cathode-material-sodium-ion-batteries"><span>Layered P<span class="hlt">2</span>-<span class="hlt">Na</span> <span class="hlt">2</span>/3 Co 1/<span class="hlt">2</span> Ti 1/<span class="hlt">2</span> O <span class="hlt">2</span> as a high-performance cathode material for sodium-ion batteries</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>Sabi, Noha; Doubaji, Siham; Hashimoto, Kazuki</p> <p></p> <p>Layered oxides are regarded as promising cathode materials for sodium-ion batteries. We present <span class="hlt">Na</span><span class="hlt">2</span>/3Co1/<span class="hlt">2</span>Ti1/<span class="hlt">2</span>O<span class="hlt">2</span> as a potential new cathode material for sodium-ion batteries. The crystal features and morphology of the pristine powder were characterized by X-ray diffraction (XRD) and scanning electron microscopy (SEM). The cathode material is evaluated in galvanostatic charge-discharge and galvanostatic intermittent titration tests, as well as ex-situ X-ray diffraction analysis. Synthesized by a high-temperature solid state reaction, <span class="hlt">Na</span><span class="hlt">2</span>/3Co1/<span class="hlt">2</span>Ti1/<span class="hlt">2</span>O<span class="hlt">2</span> crystallizes in P<span class="hlt">2</span>-type structure with P6(3)/mmc space group. The material presents reversible electrochemical behavior and delivers a specific discharge capacity of 100 mAh g(-1) when tested in <span class="hlt">Na</span> halfmore » cells between <span class="hlt">2</span>.0 and 4.<span class="hlt">2</span> V (vs. <span class="hlt">Na+/Na</span>), with capacity retention of 98% after 50 cycles. Furthermore, the electrochemical cycling of this titanium-containing material evidenced a reduction of the potential jumps recorded in the NaxCoO<span class="hlt">2</span> parent phase, revealing a positive impact of Ti substitution for Co. The ex-situ XRD measurements confirmed the reversibility and stability of the material. No structural changes were observed in the XRD patterns, and the P<span class="hlt">2</span>-type structure was stable during the charge/discharge process between <span class="hlt">2</span>.0 and 4.<span class="hlt">2</span> V vs. <span class="hlt">Na+/Na</span>. These outcomes will contribute to the progress of developing low cost electrode materials for sodium-ion batteries. (C) 2017 Elsevier B.V. All rights reserved.« 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('https://www.ncbi.nlm.nih.gov/pubmed/20232677','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/20232677"><span>Removal of Ca<span class="hlt">2</span>+ and Zn<span class="hlt">2</span>+ from aqueous solutions by zeolites <span class="hlt">Na</span>P and KP.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Yusof, Alias Mohd; Malek, Nik Ahmad Nizam Nik; Kamaruzaman, Nurul Asyikin; Adil, Muhammad</p> <p>2010-01-01</p> <p>Zeolites P in sodium (<span class="hlt">Na</span>P) and potassium (KP) forms were used as adsorbents for the removal of calcium (Ca<span class="hlt">2</span>+) and zinc (Zn<span class="hlt">2</span>+) cations from aqueous solutions. Zeolite KP was prepared by ion exchange of K+ with <span class="hlt">Na</span>+ which neutralizes the negative charge of the zeolite P framework structure. The ion exchange capacity of K+ on zeolite <span class="hlt">Na</span>P was determined through the Freundlich isotherm equilibrium study. Characterization of zeolite KP was determined using infrared spectroscopy and X-ray diffraction (XRD) techniques. From the characterization, the structure of zeolite KP was found to remain stable after the ion exchange process. Zeolites KP and <span class="hlt">Na</span>P were used for the removal of Ca and Zn from solution. The amount of Ca<span class="hlt">2</span>+ and Zn<span class="hlt">2</span>+ in aqueous solution before and after the adsorption by zeolites was analysed using the flame atomic absorption spectroscopy method. The removal of Ca<span class="hlt">2</span>+ and Zn<span class="hlt">2</span>+ followed the Freundlich isotherm rather than the Langmuir isotherm model. This result also revealed that zeolite KP adsorbs Ca<span class="hlt">2</span>+ and Zn<span class="hlt">2</span>+ more than zeolite <span class="hlt">Na</span>P and proved that modification of zeolite <span class="hlt">Na</span>P with potassium leads to an increase in the adsorption efficiency of the zeolite. Therefore, the zeolites <span class="hlt">Na</span>P and KP can be used for water softening (Ca removal) and reducing water pollution/toxicity (Zn removal).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018E%26ES..141a2023P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018E%26ES..141a2023P"><span>Absorption of CO<span class="hlt">2</span> from modified flue gases of power generation Tarahan chemically using <span class="hlt">Na</span>OH and <span class="hlt">Na</span><span class="hlt">2</span>CO3 and biologically using microalgae</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Purba, Elida; Agustina, Dewi; Putri Pertama, Finka; Senja, Fita</p> <p>2018-03-01</p> <p>This research was carried out on the absorption of CO<span class="hlt">2</span> from the modified flue gases of power generation Tarahan using <span class="hlt">Na</span>OH (sodium hydroxide) and <span class="hlt">Na</span><span class="hlt">2</span>CO3 (sodium carbonate). The operation was conducted in a packed column absorber and then the output gases from the packed column was fed into photo-bioreactor for biological absorption. In the photo-bioreactor, two species of microalgae, N. occulata and T. chuii, were cultivated to both absorb CO<span class="hlt">2</span> gas and to produce biomass for algal oil. The aims of this research were, first, to determine the effect of absorbent flow rate on the reduction of CO<span class="hlt">2</span> and on the decrease of output gas temperature, second, to determine the characteristics of methyl ester obtained from biological absorption process. Flow rates of the absorbent were varied as 1, <span class="hlt">2</span>, and 3 l/min. The concentrations of <span class="hlt">Na</span>OH and <span class="hlt">Na</span><span class="hlt">2</span>CO3 were 1 M at a constant gas flow rate of 6 l/min. The output concentrations of CO<span class="hlt">2</span> from the absorber was analyzed using Gas Chromatography 2014-AT SHIMADZU Corp 08128. The results show that both of the absorbents give different trends. From the absorption using <span class="hlt">Na</span>OH, it can be concluded that the higher the flow rate, the higher the absorption rate obtained. The highest flow rate achieved maximum absorption of 100%. On the other hand, absorption with <span class="hlt">Na</span><span class="hlt">2</span>CO3 revealed the opposite trend where the higher the flow rates the lower the absorption rate. The highest absorption using <span class="hlt">Na</span><span class="hlt">2</span>CO3 was obtained with the lowest flow rate, 1 l/min, that was 45,5%. As the effect of flow rate on output gas temperature, the temperature decreased with increasing flow rates for both absorbents. The output gas temperature for <span class="hlt">Na</span>OH and <span class="hlt">Na</span><span class="hlt">2</span>CO3 were consecutively 35 °C and 31 °C with inlet gas temperature of 50°C. Absorption of CO<span class="hlt">2</span> biologically resulted a reduction of CO<span class="hlt">2</span> up to 60% from the input gas concentration. Algal oil was extracted with mixed hexane and chloroform to obtain algal oil. Extracted oil was transesterified to methyl ester using sodium</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22274215-structure-properties-nafeo-sub-type-ternary-sodium-iridates','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22274215-structure-properties-nafeo-sub-type-ternary-sodium-iridates"><span>Structure and properties of α-<span class="hlt">Na</span>FeO{sub <span class="hlt">2</span>}-type ternary sodium iridates</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>Baroudi, Kristen, E-mail: kbaroudi@princeton.edu; Yim, Cindi; Wu, Hui</p> <p>2014-02-15</p> <p>The synthesis, structure, and elementary magnetic and electronic properties are reported for layered compounds of the type <span class="hlt">Na</span>{sub 3−x}MIr{sub <span class="hlt">2</span>}O{sub 6} and <span class="hlt">Na</span>{sub 3−x}M{sub <span class="hlt">2</span>}IrO{sub 6}, where M is a transition metal from the 3d series (M=Zn, Cu, Ni, Co, Fe and Mn). The rhombohedral structures, in space group R−3m, were determined by refinement of neutron and synchrotron powder diffraction data. No clear evidence for long range <span class="hlt">2</span>:1 or 1:<span class="hlt">2</span> honeycomb-like M/Ir ordering was found in the neutron powder diffraction patterns except in the case of M=Zn, and thus in general the compounds are best designated as sodium deficient α-<span class="hlt">Na</span>FeO{submore » <span class="hlt">2</span>}-type phases with formulas <span class="hlt">Na</span>{sub 1−x}M{sub 1/3}Ir{sub <span class="hlt">2</span>/3}O{sub <span class="hlt">2</span>} or <span class="hlt">Na</span>{sub 1−x}M{sub <span class="hlt">2</span>/3}Ir{sub 1/3}O{sub <span class="hlt">2</span>}. Synchrotron powder diffraction patterns indicate that several of the compounds likely have honeycomb in-plane metal–iridium ordering with disordered stacking of the layers. All the compounds are sodium deficient under our synthetic conditions and are black and insulating. Weiss constants derived from magnetic susceptibility measurements indicate that <span class="hlt">Na</span>{sub 0.62}Mn{sub 0.61}Ir{sub 0.39}O{sub <span class="hlt">2</span>}, <span class="hlt">Na</span>{sub 0.80}Fe{sub <span class="hlt">2</span>/3}Ir{sub 1/3}O{sub <span class="hlt">2</span>}, <span class="hlt">Na</span>{sub 0.92}Ni{sub 1/3}Ir{sub <span class="hlt">2</span>/3}O{sub <span class="hlt">2</span>}, <span class="hlt">Na</span>{sub 0.86}Cu{sub 1/3}Ir{sub <span class="hlt">2</span>/3}O{sub <span class="hlt">2</span>}, and <span class="hlt">Na</span>{sub 0.89}Zn{sub 1/3}Ir{sub <span class="hlt">2</span>/3}O{sub <span class="hlt">2</span>} display dominant antiferromagnetic interactions. For <span class="hlt">Na</span>{sub 0.90}Co{sub 1/3}Ir{sub <span class="hlt">2</span>/3}O{sub <span class="hlt">2</span>} the dominant magnetic interactions at low temperature are ferromagnetic while at high temperatures they are antiferromagnetic; there is also a change in the effective moment. Low temperature specific heat measurements (to <span class="hlt">2</span> K) on <span class="hlt">Na</span>{sub 0.92}Ni{sub 1/3}Ir{sub <span class="hlt">2</span>/3}O{sub <span class="hlt">2</span>} indicate the presence of a broad magnetic ordering transition. X-ray absorption spectroscopy shows that iridium is at or close to the 4+ oxidation state in all compounds. {sup 23}<span class="hlt">Na</span> nuclear magnetic resonance measurements comparing <span class="hlt">Na</span>{sub <span class="hlt">2</span>}IrO{sub 3} to <span class="hlt">Na</span>{sub 0.92}Ni{sub 1</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1983JNuM..119..296S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1983JNuM..119..296S"><span>Threshold oxygen levels in <span class="hlt">Na</span>(I) for the formation of <span class="hlt">Na</span>CrO <span class="hlt">2</span>(s) on 18-8 stainless steels from accurate thermodynamic measurements</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sreedharan, O. M.; Madan, B. S.; Gnanamoorthy, J. B.</p> <p>1983-12-01</p> <p>The compound <span class="hlt">Na</span>CrO <span class="hlt">2</span>(s) is an important corrosion product in sodium-cooled LMFBRs. The standard Gibbs energy of formation of <span class="hlt">Na</span>CrO <span class="hlt">2</span>(s) is required for the computation of threshold oxygen levels in <span class="hlt">Na</span>(1) for the formation of <span class="hlt">Na</span>CrO <span class="hlt">2</span>(s) on 18-8 stainless steels. For this purpose the emf of the galvanic cell: Pt, <span class="hlt">Na</span>CrO <span class="hlt">2</span>, Cr <span class="hlt">2</span>O 3, <span class="hlt">Na</span> <span class="hlt">2</span>CrO 4/15 YSZ/O <span class="hlt">2</span> ( P O <span class="hlt">2</span> = 0.21 atm, air), Pt was measured over 784-1012 K to be: (E±4.4)(mV) = 483.67-0.34155 T(K). From this, the standard Gibbs energy of formation of <span class="hlt">Na</span>CrO <span class="hlt">2</span>(s) from the elements ( ΔG f,T0) and from the oxides ( ΔG f,OX,T0) was calculated to be: [ΔG f,T0(<span class="hlt">Na</span>CrO <span class="hlt">2</span>, s)±1.86] (kJ/mol) =-869.98 + 0.18575 T(K) , [ΔG f,OX,T0(<span class="hlt">Na</span>Cr0 <span class="hlt">2</span>, s)±4.8] (kJ/mol) = -104.25-0.00856 T(K) . The molar heat capacity, C P0, of <span class="hlt">Na</span>CrO <span class="hlt">2</span>(s) was measured by DSC to be (350-600 K): C P0(<span class="hlt">Na</span>CrO <span class="hlt">2</span>, s) (J/K mol) = 27.15 + 0.1247 T (K) , From these data, values of -99.3 kJ/mol and 91.6 J/K mol were computed for ΔH f,2980 and S 2980 of <span class="hlt">Na</span>CrO <span class="hlt">2</span>(s). The internal consistency was checked with the use of enthalpy data on <span class="hlt">Na</span> <span class="hlt">2</span>CrO 4(s). From the standard Gibbs energy of formation of <span class="hlt">Na</span>CrO <span class="hlt">2</span>(s) the equation logC 0(wppm) = 3.9905-3147.6 T(K) was derived, where C 0 is the threshold oxygen level for the formation of <span class="hlt">Na</span>CrO <span class="hlt">2</span>(s) on 18-8 stainless steels.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28683437','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28683437"><span>Role of <span class="hlt">Na</span>+/Ca<span class="hlt">2</span>+ Exchangers in Therapy Resistance of Medulloblastoma Cells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Pelzl, Lisann; Hosseinzadeh, Zohreh; Al-Maghout, Tamer; Singh, Yogesh; Sahu, Itishri; Bissinger, Rosi; Schmidt, Sebastian; Alkahtani, Saad; Stournaras, Christos; Toulany, Mahmoud; Lang, Florian</p> <p>2017-01-01</p> <p>Alterations of cytosolic Ca<span class="hlt">2</span>+-activity ([Ca<span class="hlt">2</span>+]i) are decisive in the regulation of tumor cell proliferation, migration and survival. Transport processes participating in the regulation of [Ca<span class="hlt">2</span>+]i include Ca<span class="hlt">2</span>+ extrusion through K+-independent (NCX) and/or K+-dependent (NCKX) <span class="hlt">Na</span>+/Ca<span class="hlt">2</span>+-exchangers. The present study thus explored whether medulloblastoma cells express <span class="hlt">Na</span>+/Ca<span class="hlt">2</span>+-exchangers, whether expression differs between therapy sensitive D283 and therapy resistant UW228-3 medulloblastoma cells, and whether <span class="hlt">Na</span>+/Ca<span class="hlt">2</span>+-exchangers participate in the regulation of cell survival. In therapy sensitive D283 and therapy resistant UW228-3 medulloblastoma cells transcript levels were estimated by RT-PCR, protein abundance by Western blotting, cytosolic Ca<span class="hlt">2</span>+-activity ([Ca<span class="hlt">2</span>+]i) from Fura-<span class="hlt">2</span>-fluorescence, <span class="hlt">Na</span>+/ Ca<span class="hlt">2</span>+-exchanger activity from the increase of [Ca<span class="hlt">2</span>+]i (Δ[Ca<span class="hlt">2</span>+]i) and from whole cell current (Ica) following abrupt replacement of <span class="hlt">Na</span>+ containing (130 mM) and Ca<span class="hlt">2</span>+ free by <span class="hlt">Na</span>+ free and Ca<span class="hlt">2</span>+ containing (<span class="hlt">2</span> mM) extracellular perfusate as well as cell death from PI -staining and annexin-V binding in flow cytometry. The transcript levels of NCX3, NCKX<span class="hlt">2</span>, and NCKX5, protein abundance of NCX3, slope and peak of Δ[Ca<span class="hlt">2</span>+]i as well as Ica were significantly lower in therapy sensitive D283 than in therapy resistant UW228-3 medulloblastoma cells. The <span class="hlt">Na</span>+/Ca<span class="hlt">2</span>+-exchanger inhibitor KB-R7943 (10 µM) significantly blunted Δ[Ca<span class="hlt">2</span>+]i, and augmented the ionizing radiation-induced apoptosis but did not significantly modify clonogenicity of medulloblastoma cells. Apoptosis was further enhanced by NCX3 silencing. <span class="hlt">Na</span>+/Ca<span class="hlt">2</span>+-exchanger activity significantly counteracts apoptosis but does not significantly affect clonogenicity after radiation of medulloblastoma cells. © 2017 The Author(s). Published by S. Karger AG, Basel.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012REDS..167..807W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012REDS..167..807W"><span>Luminescence characteristics of Dy3+ activated <span class="hlt">Na</span> <span class="hlt">2</span>Sr <span class="hlt">2</span>Mg (BO 3)<span class="hlt">2</span>F <span class="hlt">2</span>: Dy 3+ phosphor</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wani, Javaid A.; Dhoble, N. S.; Dhoble, S. J.</p> <p>2012-11-01</p> <p>In this paper, we have reported a new <span class="hlt">Na</span> <span class="hlt">2</span>Sr <span class="hlt">2</span>Mg (BO 3)<span class="hlt">2</span>F <span class="hlt">2</span>:Dy 3+ thermoluminescence (TL) phosphor prepared via the wet chemical method. Prepared phosphor was characterized by X-ray powder diffraction, photoluminescence (PL), TL and scanning electronmicroscopy techniques. The scanning electronmicroscopic image of <span class="hlt">Na</span> <span class="hlt">2</span>Sr <span class="hlt">2</span>Mg (BO 3)<span class="hlt">2</span>F <span class="hlt">2</span>:Dy 3+ phosphor confirms the micron size of particles. Under the PL study, the characteristic emission spectrum of Dy 3+ corresponding to 4F 9/<span class="hlt">2</span>→6H 15/<span class="hlt">2</span> (481 nm) and 4F 9/<span class="hlt">2</span>→6H 13/<span class="hlt">2</span> (576 nm) transitions was observed. The TL property of the as prepared phosphor was also found to be good. TL intensity of <span class="hlt">Na</span> <span class="hlt">2</span>Sr<span class="hlt">2</span>Mg(BO 3)F <span class="hlt">2</span>:Dy 3+ phosphors at 0.99 kGy exposure of γ-irradiations was compared with standard CaSO 4:Dy phosphor. It was seen that TL intensity of <span class="hlt">Na</span> <span class="hlt">2</span>Sr <span class="hlt">2</span>Mg (BO 3)<span class="hlt">2</span>F <span class="hlt">2</span>: Dy 3+ phosphors is 1.1 times less compared with the standard CaSO 4:Dy TL dosimeter phosphor. The kinetic parameters are also discussed in detail. The values of activation energy E (eV) and frequency factor S (s -1) were found to be 0.57 eV and 1.25×106 s-1, respectively.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25644099','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25644099"><span>Investigating the influence of <span class="hlt">Na</span>+ and Sr<span class="hlt">2</span>+ on the structure and solubility of SiO<span class="hlt">2</span>-TiO<span class="hlt">2</span>-CaO-<span class="hlt">Na</span><span class="hlt">2</span>O/SrO bioactive glass.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Li, Y; Placek, L M; Coughlan, A; Laffir, F R; Pradhan, D; Mellott, N P; Wren, A W</p> <p>2015-02-01</p> <p>This study was conducted to determine the influence that network modifiers, sodium (<span class="hlt">Na</span>+) and strontium (Sr<span class="hlt">2</span>+), have on the solubility of a SiO<span class="hlt">2</span>-TiO<span class="hlt">2</span>-CaO-<span class="hlt">Na</span><span class="hlt">2</span>O/SrO bioactive glass. Glass characterization determined each composition had a similar structure, i.e. bridging to non-bridging oxygen ratio determined by X-ray photoelectron spectroscopy. Magic angle spinning nuclear magnetic resonance (MAS-NMR) confirmed structural similarities as each glass presented spectral shifts between -84 and -85 ppm. Differential thermal analysis and hardness testing revealed higher glass transition temperatures (Tg 591-760 °C) and hardness values (<span class="hlt">2</span>.4-6.1 GPa) for the Sr<span class="hlt">2</span>+ containing glasses. Additionally the Sr<span class="hlt">2</span>+ (~250 mg/L) containing glasses displayed much lower ion release rates than the <span class="hlt">Na</span>+ (~1,200 mg/L) containing glass analogues. With the reduction in ion release there was an associated reduction in solution pH. Cytotoxicity and cell adhesion studies were conducted using MC3T3 Osteoblasts. Each glass did not significantly reduce cell numbers and osteoblasts were found to adhere to each glass surface.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/1043784-characterization-photochemical-processes-h2-production-cds-nanorod-fefe-hydrogenase-complexes','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/1043784-characterization-photochemical-processes-h2-production-cds-nanorod-fefe-hydrogenase-complexes"><span>Characterization of Photochemical Processes for H<span class="hlt">2</span> Production by CdS Nanorod-[FeFe] Hydrogenase Complexes</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>Brown, K. A.; Wilker, M. B.; Boehm, M.</p> <p>2012-03-28</p> <p>We have developed complexes of CdS nanorods capped with 3-mercaptopropionic acid (MPA) and Clostridium acetobutylicum [FeFe]-hydrogenase I (<span class="hlt">CaI</span>) that photocatalyze reduction of H{sup +} to H{sub <span class="hlt">2</span>} at a <span class="hlt">CaI</span> turnover frequency of 380-900 s{sup -1} and photon conversion efficiencies of up to 20% under illumination at 405 nm. In this paper, we focus on the compositional and mechanistic aspects of CdS:CaI complexes that control the photochemical conversion of solar energy into H{sub <span class="hlt">2</span>}. Self-assembly of CdS with <span class="hlt">CaI</span> was driven by electrostatics, demonstrated as the inhibition of ferredoxin-mediated H{sub <span class="hlt">2</span>} evolution by <span class="hlt">CaI</span>. Production of H{sub <span class="hlt">2</span>} by CdS:CaImore » was observed only under illumination and only in the presence of a sacrificial donor. We explored the effects of the CdS:CaI molar ratio, sacrificial donor concentration, and light intensity on photocatalytic H{sub <span class="hlt">2</span>} production, which were interpreted on the basis of contributions to electron transfer, hole transfer, or rate of photon absorption, respectively. Each parameter was found to have pronounced effects on the CdS:CaI photocatalytic activity. Specifically, we found that under 405 nm light at an intensity equivalent to total AM 1.5 solar flux, H{sub <span class="hlt">2</span>} production was limited by the rate of photon absorption ({approx}1 ms{sup -1}) and not by the turnover of <span class="hlt">CaI</span>. Complexes were capable of H{sub <span class="hlt">2</span>} production for up to 4 h with a total turnover number of 106 before photocatalytic activity was lost. This loss correlated with inactivation of <span class="hlt">CaI</span>, resulting from the photo-oxidation of the CdS capping ligand MPA.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70010037','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70010037"><span>Borate mineral assemblages in the system <span class="hlt">Na</span><span class="hlt">2</span>OCaOMgOB<span class="hlt">2</span>O3H<span class="hlt">2</span>O</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Christ, C.L.; Truesdell, A.H.; Erd, Richard C.</p> <p>1967-01-01</p> <p>he significant known hydrated borate mineral assemblages (principally of the western United States) in the system <span class="hlt">Na</span><span class="hlt">2</span>OCaOz.sbnd;MgOB<span class="hlt">2</span>O3H<span class="hlt">2</span>O are expressible in three ternary composition diagrams. Phase rule interpretation of the diagrams is consistent with observation, if the activity of H<span class="hlt">2</span>O is generally considered to be determined by the geologic environment. The absence of conflicting tie-lines on a diagram indicates that the several mineral assemblages of the diagram were formed under relatively narrow ranges of temperature and pressure. The known structural as well as empirical formulas for the minerals are listed, and the more recent (since 1960) crystal structure findings are discussed briefly. Schematic Gibbs free energy-composition diagrams based on known solubility-temperature relations in the systems <span class="hlt">Na</span><span class="hlt">2</span>B4O7-H<span class="hlt">2</span>O and <span class="hlt">Na</span><span class="hlt">2</span>B4O7-<span class="hlt">Na</span>Cl-H<span class="hlt">2</span>O, are highly useful in the interpretation and prediction of the stability relations in these systems; in particular these diagrams indicate clearly that tincalconite, although geologically important, is everywhere a metastable phase. Crystal-chemical considerations indicate that the same thermodynamic and kinetic behavior observed in the <span class="hlt">Na</span><span class="hlt">2</span>B4O7-H<span class="hlt">2</span>O system will hold in the Ca<span class="hlt">2</span>B6O11-H<span class="hlt">2</span>O system. This conclusion is confirmed by the petrologic evidence. The chemical relations among the mineral assemblages of a ternary diagram are expressed by a schematic "activity-activity" diagram. These activity-activity diagrams permit the tracing-out of the paragenetic sequences as a function of changing cation and H<span class="hlt">2</span>O activities. ?? 1967.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22927545','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22927545"><span><span class="hlt">Na</span>+,K+-ATPase functionally interacts with the plasma membrane <span class="hlt">Na</span>+,Ca<span class="hlt">2</span>+ exchanger to prevent Ca<span class="hlt">2</span>+ overload and neuronal apoptosis in excitotoxic stress.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Sibarov, Dmitry A; Bolshakov, Artemiy E; Abushik, Polina A; Krivoi, Igor I; Antonov, Sergei M</p> <p>2012-12-01</p> <p>Using a fluorescent viability assay, immunocytochemistry, patch-clamp recordings, and Ca(<span class="hlt">2</span>+) imaging analysis, we report that ouabain, a specific ligand of the <span class="hlt">Na</span>(+),K(+)-ATPase cardiac glycoside binding site, can prevent glutamate receptor agonist-induced apoptosis in cultured rat cortical neurons. In our model of excitotoxicity, a 240-min exposure to 30 μM N-methyl-d-aspartate (NMDA) or kainate caused apoptosis in ∼50% of neurons. These effects were accompanied by a significant decrease in the number of neurons that were immunopositive for the antiapoptotic peptide Bcl-<span class="hlt">2</span>. Apoptotic injury was completely prevented when the agonists were applied together with 0.1 or 1 nM ouabain, resulting in a greater survival of neurons, and the percentage of neurons expressing Bcl-<span class="hlt">2</span> remained similar to those obtained without agonist treatments. In addition, subnanomolar concentrations of ouabain prevented the increase of spontaneous excitatory postsynaptic current's frequency and the intracellular Ca(<span class="hlt">2</span>+) overload induced by excitotoxic insults. Loading neurons with 1,<span class="hlt">2</span>-bis(<span class="hlt">2</span>-aminophenoxy)ethane-N,N,N',N'-tetraacetic acid or inhibition of the plasma membrane <span class="hlt">Na</span>(+),Ca(<span class="hlt">2</span>+)-exchanger by <span class="hlt">2-(2</span>-(4-(4-nitrobenzyloxy)phenyl)ethyl)isothiourea methanesulfonate (KB-R7943) eliminated ouabain's effects on NMDA- or kainite-evoked enhancement of spontaneous synaptic activity. Our data suggest that during excitotoxic insults ouabain accelerates Ca(<span class="hlt">2</span>+) extrusion from neurons via the <span class="hlt">Na</span>(+),Ca(<span class="hlt">2</span>+) exchanger. Because intracellular Ca(<span class="hlt">2</span>+) accumulation caused by the activation of glutamate receptors and boosted synaptic activity represents a key factor in triggering neuronal apoptosis, up-regulation of Ca(<span class="hlt">2</span>+) extrusion abolishes its development. These antiapoptotic effects are independent of <span class="hlt">Na</span>(+),K(+)-ATPase ion transport function and are initiated by concentrations of ouabain that are within the range of an endogenous analog, suggesting a novel functional role for <span class="hlt">Na</span>(+),K(+)-ATPase in</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70018221','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70018221"><span>Calculation of the vapor-saturated liquidus for the <span class="hlt">Na</span>Cl-CO<span class="hlt">2</span>-H<span class="hlt">2</span>O system</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Barton, P.B.; I-Ming, C.</p> <p>1993-01-01</p> <p>The polybaric liquidus surface for the H<span class="hlt">2</span>O-rich corner of the <span class="hlt">Na</span>Cl-CO<span class="hlt">2</span>-H<span class="hlt">2</span>O ternary is calculated, relying heavily on 1. (1) a Henry's law equation for CO<span class="hlt">2</span> in brines (modified from Drummond, 1981), <span class="hlt">2</span>. (<span class="hlt">2</span>) the assumption that the contributions of dissolved <span class="hlt">Na</span>Cl and CO<span class="hlt">2</span> in lowering the activity of H<span class="hlt">2</span>O are additive, and 3. (3) data on the CO<span class="hlt">2</span> clathrate solid solution (nominally CO<span class="hlt">2</span> ?? 7.3H<span class="hlt">2</span>O, but ranging from 5.75 to 8 or 9 H<span class="hlt">2</span>O) from Bozzo et al. (1975). The variation with composition of the activity of CO<span class="hlt">2</span>??7.3H<span class="hlt">2</span>O, or any other composition within the clathrate field, is small, thereby simplifying the calculations appreciably. Ternary invariant points are 1. (1) ternary eutectic at -21.5??C, with ice + clathrate + hydrohalite <span class="hlt">Na</span>Cl-??H<span class="hlt">2</span>O + brine m<span class="hlt">Na</span>Cl = 5.15, mco<span class="hlt">2</span> = 0.22 + vapor Ptotal ??? Pco<span class="hlt">2</span> = 5.7 atm; <span class="hlt">2</span>. (<span class="hlt">2</span>) peritectic at -9.6??C, with clathrate + hydrohalite + liquid CO<span class="hlt">2</span> + brine m<span class="hlt">Na</span>Cl = 5.18, mco<span class="hlt">2</span> = 0.55 + vapor (Ptotal ??? Pco<span class="hlt">2</span> = 26.47 atm); and 3. (3) peritectic slightly below +0.1 ??C, with halite + hydrohalite + liquid CO<span class="hlt">2</span> + brine (m<span class="hlt">Na</span>Cl ??? 5.5, mco<span class="hlt">2</span> ??? 0.64) + vapor (Ptotal ??? Pco<span class="hlt">2</span> ??? 34 atm). CO<span class="hlt">2</span> isobars have been contoured on the ternary liquidus and also on the 25??C isotherm. An important caveat regarding the application of this information to the interpretation of the freezing-thawing behavior of fluid inclusions is that metastable behavior is a common characteristic of the clathrate. ?? 1993.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1335455','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1335455"><span>CO Oxidation and Subsequent CO <span class="hlt">2</span> Chemisorption on Alkaline Zirconates: Li <span class="hlt">2</span> ZrO 3 and <span class="hlt">Na</span> <span class="hlt">2</span> ZrO 3</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>Alcántar-Vázquez, Brenda; Duan, Yuhua; Pfeiffer, Heriberto</p> <p></p> <p>Here, two different alkaline zirconates (Li <span class="hlt">2</span>ZrO 3 and <span class="hlt">Na</span> <span class="hlt">2</span>ZrO 3) were studied as possible bifunctional catalytic-captor materials for CO oxidation and the subsequent CO <span class="hlt">2</span> chemisorption process. Initially, CO oxidation reactions were analyzed in a catalytic reactor coupled to a gas chromatograph, using Li <span class="hlt">2</span>ZrO 3 and <span class="hlt">Na</span> <span class="hlt">2</span>ZrO 3, under different O <span class="hlt">2</span> partial flows. We found results clearly showed that <span class="hlt">Na</span> <span class="hlt">2</span>ZrO 3 possesses much better catalytic properties than Li <span class="hlt">2</span>ZrO 3. After the CO-O <span class="hlt">2</span> oxidation catalytic analysis, CO<span class="hlt">2</span> chemisorption process was analyzed by thermogravimetric analysis, only for the <span class="hlt">Na</span> <span class="hlt">2</span>ZrO 3 ceramic. The resultsmore » confirmed that <span class="hlt">Na</span> <span class="hlt">2</span>ZrO 3 is able to work as a bifunctional material (CO oxidation and subsequent CO <span class="hlt">2</span> chemisorption), although the kinetic CO <span class="hlt">2</span> capture process was not the best one under the physicochemical condition used in this case. For <span class="hlt">Na</span> <span class="hlt">2</span>ZrO 3, the best CO conversions were found between 445 and 580 °C (100%), while Li <span class="hlt">2</span>ZrO 3 only showed a 35% of efficiency between 460 and 503 °C. However, in the <span class="hlt">Na</span> <span class="hlt">2</span>ZrO 3 case, at temperatures higher than 580 °C its catalytic activity gradually decreases as a result of CO <span class="hlt">2</span> capture process. Finally, all these experiments were compared and supported with theoretical thermodynamic data.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1335455-co-oxidation-subsequent-co2-chemisorption-alkaline-zirconates-li2-zro3-na2-zro3','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1335455-co-oxidation-subsequent-co2-chemisorption-alkaline-zirconates-li2-zro3-na2-zro3"><span>CO Oxidation and Subsequent CO <span class="hlt">2</span> Chemisorption on Alkaline Zirconates: Li <span class="hlt">2</span> ZrO 3 and <span class="hlt">Na</span> <span class="hlt">2</span> ZrO 3</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Alcántar-Vázquez, Brenda; Duan, Yuhua; Pfeiffer, Heriberto</p> <p>2016-08-26</p> <p>Here, two different alkaline zirconates (Li <span class="hlt">2</span>ZrO 3 and <span class="hlt">Na</span> <span class="hlt">2</span>ZrO 3) were studied as possible bifunctional catalytic-captor materials for CO oxidation and the subsequent CO <span class="hlt">2</span> chemisorption process. Initially, CO oxidation reactions were analyzed in a catalytic reactor coupled to a gas chromatograph, using Li <span class="hlt">2</span>ZrO 3 and <span class="hlt">Na</span> <span class="hlt">2</span>ZrO 3, under different O <span class="hlt">2</span> partial flows. We found results clearly showed that <span class="hlt">Na</span> <span class="hlt">2</span>ZrO 3 possesses much better catalytic properties than Li <span class="hlt">2</span>ZrO 3. After the CO-O <span class="hlt">2</span> oxidation catalytic analysis, CO<span class="hlt">2</span> chemisorption process was analyzed by thermogravimetric analysis, only for the <span class="hlt">Na</span> <span class="hlt">2</span>ZrO 3 ceramic. The resultsmore » confirmed that <span class="hlt">Na</span> <span class="hlt">2</span>ZrO 3 is able to work as a bifunctional material (CO oxidation and subsequent CO <span class="hlt">2</span> chemisorption), although the kinetic CO <span class="hlt">2</span> capture process was not the best one under the physicochemical condition used in this case. For <span class="hlt">Na</span> <span class="hlt">2</span>ZrO 3, the best CO conversions were found between 445 and 580 °C (100%), while Li <span class="hlt">2</span>ZrO 3 only showed a 35% of efficiency between 460 and 503 °C. However, in the <span class="hlt">Na</span> <span class="hlt">2</span>ZrO 3 case, at temperatures higher than 580 °C its catalytic activity gradually decreases as a result of CO <span class="hlt">2</span> capture process. Finally, all these experiments were compared and supported with theoretical thermodynamic data.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017MNRAS.472..361R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017MNRAS.472..361R"><span>The puzzling interpretation of NIR indices: The case of <span class="hlt">Na</span>I<span class="hlt">2</span>.21</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Röck, B.; Vazdekis, A.; La Barbera, F.; Peletier, R. F.; Knapen, J. H.; Allende-Prieto, C.; Aguado, D. S.</p> <p>2017-11-01</p> <p>We present a detailed study of the <span class="hlt">Na</span> I line strength index centred in the K band at 22 100 Å (<span class="hlt">Na</span>I<span class="hlt">2</span>.21 hereafter) relying on different samples of early-type galaxies. Consistent with previous studies, we find that the observed line strength indices cannot be fit by state-of-the-art scaled-solar stellar population models, even using our newly developed models in the near infrared (NIR). The models clearly underestimate the large <span class="hlt">Na</span>I<span class="hlt">2</span>.21 values measured for most early-type galaxies. However, we develop an <span class="hlt">Na</span>-enhanced version of our newly developed models in the NIR, which - together with the effect of a bottom-heavy initial mass function - yield <span class="hlt">Na</span>I<span class="hlt">2</span>.21 indices in the range of the observations. Therefore, we suggest a scenario in which the combined effect of [<span class="hlt">Na</span>/Fe] enhancement and a bottom-heavy initial mass function are mainly responsible for the large <span class="hlt">Na</span>I<span class="hlt">2</span>.21 indices observed for most early-type galaxies. To a smaller extent, also [C/Fe] enhancement might contribute to the large observed <span class="hlt">Na</span>I<span class="hlt">2</span>.21 values.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5065245','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5065245"><span>Neuronal <span class="hlt">Na</span>+ Channels Are Integral Components of Pro-arrhythmic <span class="hlt">Na</span>+/Ca<span class="hlt">2</span>+ Signaling Nanodomain That Promotes Cardiac Arrhythmias During β-adrenergic Stimulation</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Radwański, Przemysław B.; Ho, Hsiang-Ting; Veeraraghavan, Rengasayee; Brunello, Lucia; Liu, Bin; Belevych, Andriy E.; Unudurthi, Sathya D.; Makara, Michael A.; Priori, Silvia G.; Volpe, Pompeo; Armoundas, Antonis A.; Dillmann, Wolfgang H.; Knollmann, Bjorn C.; Mohler, Peter J.; Hund, Thomas J.; Györke, Sándor</p> <p>2016-01-01</p> <p>Background Cardiac arrhythmias are a leading cause of death in the US. Vast majority of these arrhythmias including catecholaminergic polymorphic ventricular tachycardia (CPVT) are associated with increased levels of circulating catecholamines and involve abnormal impulse formation secondary to aberrant Ca<span class="hlt">2</span>+ and <span class="hlt">Na</span>+ handling. However, the mechanistic link between β-AR stimulation and the subcellular/molecular arrhythmogenic trigger(s) remains elusive. Methods and Results We performed functional and structural studies to assess Ca<span class="hlt">2</span>+ and <span class="hlt">Na</span>+ signaling in ventricular myocyte as well as surface electrocardiograms in mouse models of cardiac calsequestrin (CASQ<span class="hlt">2</span>)-associated CPVT. We demonstrate that a subpopulation of <span class="hlt">Na</span>+ channels (neuronal <span class="hlt">Na</span>+ channels; nNav) that colocalize with RyR<span class="hlt">2</span> and <span class="hlt">Na</span>+/Ca<span class="hlt">2</span>+ exchanger (NCX) are a part of the β-AR-mediated arrhythmogenic process. Specifically, augmented <span class="hlt">Na</span>+ entry via nNav in the settings of genetic defects within the RyR<span class="hlt">2</span> complex and enhanced sarcoplasmic reticulum (SR) Ca<span class="hlt">2</span>+-ATPase (SERCA)-mediated SR Ca<span class="hlt">2</span>+ refill is both an essential and a necessary factor for the arrhythmogenesis. Furthermore, we show that augmentation of <span class="hlt">Na</span>+ entry involves β-AR-mediated activation of CAMKII subsequently leading to nNav augmentation. Importantly, selective pharmacological inhibition as well as silencing of Nav1.6 inhibit myocyte arrhythmic potential and prevent arrhythmias in vivo. Conclusion These data suggest that the arrhythmogenic alteration in <span class="hlt">Na</span>+/Ca<span class="hlt">2</span>+ handling evidenced ruing β-AR stimulation results, at least in part, from enhanced <span class="hlt">Na</span>+ influx through nNav. Therefore, selective inhibition of these channels and Nav1.6 in particular can serve as a potential antiarrhythmic therapy. PMID:27747307</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017MMTB...48.1134Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017MMTB...48.1134Z"><span>Influences of <span class="hlt">Na</span><span class="hlt">2</span>O and K<span class="hlt">2</span>O Additions on Electrical Conductivity of CaO-MgO-Al<span class="hlt">2</span>O3-SiO<span class="hlt">2</span> Melts</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zhang, Guo-Hua; Zheng, Wei-Wei; Chou, Kuo-Chih</p> <p>2017-04-01</p> <p>The present study investigated the influences of <span class="hlt">Na</span><span class="hlt">2</span>O and K<span class="hlt">2</span>O additions on electrical conductivity of blast furnace type CaO-MgO-Al<span class="hlt">2</span>O3-SiO<span class="hlt">2</span> melts by the four-electrode method. Both the single addition of <span class="hlt">Na</span><span class="hlt">2</span>O or K<span class="hlt">2</span>O and the double additions of <span class="hlt">Na</span><span class="hlt">2</span>O and K<span class="hlt">2</span>O were studied. It was found that electrical conductivity monotonously increased as the amount of <span class="hlt">Na</span><span class="hlt">2</span>O addition was gradually increased, whereas, when K<span class="hlt">2</span>O was added, there was a continuous decrease of electrical conductivity. With melts containing both <span class="hlt">Na</span><span class="hlt">2</span>O and K<span class="hlt">2</span>O, electrical conductivity first decreased but then increased when <span class="hlt">Na</span><span class="hlt">2</span>O was gradually substituted for K<span class="hlt">2</span>O while keeping the molar fractions of other components constant. In other words, the mixed-alkali effect took place in CaO-Mg-Al<span class="hlt">2</span>O3-SiO<span class="hlt">2</span>-ΣR<span class="hlt">2</span>O melts.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://eric.ed.gov/?q=%22principles+of+macroeconomics%22&pg=2&id=ED150057','ERIC'); return false;" href="https://eric.ed.gov/?q=%22principles+of+macroeconomics%22&pg=2&id=ED150057"><span>An Evaluation of the Cognitive and Affective Performance of an Integrated Set of <span class="hlt">CAI</span> Materials in the Principles of Macroeconomics. Studies in Economic Education, No. 4.</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>Daellenbach, Lawrence A.; And Others</p> <p></p> <p>The purpose of this study was to determine the effect of computer assisted instruction (<span class="hlt">CAI</span>) on the cognitive and affective development of college students enrolled in a principles of macroeconomics course. The hypotheses of the experiment were stated as follows: In relation to the traditional principles course, the experimental treatment will…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/17676823','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/17676823"><span>Crystallization kinetics of bioactive glasses in the ZnO-<span class="hlt">Na</span><span class="hlt">2</span>O-CaO-SiO<span class="hlt">2</span> system.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Malavasi, Gianluca; Lusvardi, Gigliola; Pedone, Alfonso; Menziani, Maria Cristina; Dappiaggi, Monica; Gualtieri, Alessandro; Menabue, Ledi</p> <p>2007-08-30</p> <p>The crystallization kinetics of <span class="hlt">Na</span>(<span class="hlt">2</span>)O.CaO.<span class="hlt">2</span>SiO(<span class="hlt">2</span>) (x = 0) and 0.68ZnO.<span class="hlt">Na</span>(<span class="hlt">2</span>)O.CaO.<span class="hlt">2</span>SiO(<span class="hlt">2</span>) (x = 0.68, where x is the ZnO stoichiometric coefficient in the glass formula) bioactive glasses have been studied using both nonisothermal and isothermal methods. The results obtained from isothermal XRPD analyses have showed that the first glass crystallizes into the isochemical <span class="hlt">Na</span>(<span class="hlt">2</span>)CaSi(<span class="hlt">2</span>)O(6) phase, whereas the <span class="hlt">Na</span>(<span class="hlt">2</span>)ZnSiO(4) crystalline phase is obtained from the Zn-rich glass, in addition to <span class="hlt">Na</span>(<span class="hlt">2</span>)CaSi(<span class="hlt">2</span>)O(6). The activation energy (Ea) for the crystallization of the <span class="hlt">Na</span>(<span class="hlt">2</span>)O.CaO.<span class="hlt">2</span>SiO(<span class="hlt">2</span>) glass is 193 +/- 10 and 203 +/- 5 kJ/mol from the isothermal in situ XRPD and nonisothermal DSC experiments, respectively. The Avrami exponent n determined from the isothermal method is 1 at low temperature (530 degrees C), and its value increases linearly with temperature increase up to <span class="hlt">2</span> at 607 degrees C. For the crystallization of <span class="hlt">Na</span>(<span class="hlt">2</span>)CaSi(<span class="hlt">2</span>)O(6) from the Zn-containing glass, higher values of both the crystallization temperature (667 and 661 degrees C) and Ea (223 +/- 10 and 211 +/- 5 kJ/mol) have been found from the isothermal and nonisothermal methods, respectively. The <span class="hlt">Na</span>(<span class="hlt">2</span>)ZnSiO(4) crystalline phase crystallizes at lower temperature with respect to <span class="hlt">Na</span>(<span class="hlt">2</span>)CaSi(<span class="hlt">2</span>)O(6), and the Ea value is 266 +/- 20 and 245 +/- 15 kJ/mol from the isothermal and nonisothermal methods, respectively. The results of this work show that the addition of Zn favors the crystallization from the glass at lower temperature with respect to the Zn-free glass. In fact, it causes an increase of Ea for the <span class="hlt">Na</span> diffusion process, determined using MD simulations, and consequently an overall increase of Ea for the crystallization process of <span class="hlt">Na</span>(<span class="hlt">2</span>)CaSi(<span class="hlt">2</span>)O(6). Our results show good agreement between the Ea and n values obtained with the two different methods and confirm the reliability of the nonisothermal method applied to kinetic crystallization of glassy systems. This study allows the determination of the temperature</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4463720','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4463720"><span>Optimization of <span class="hlt">Na</span>OH Molarity, LUSI Mud/Alkaline Activator, and <span class="hlt">Na</span><span class="hlt">2</span>SiO3/<span class="hlt">Na</span>OH Ratio to Produce Lightweight Aggregate-Based Geopolymer</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Abdul Razak, Rafiza; Abdullah, Mohd Mustafa Al Bakri; Hussin, Kamarudin; Ismail, Khairul Nizar; Hardjito, Djwantoro; Yahya, Zarina</p> <p>2015-01-01</p> <p>This paper presents the mechanical function and characterization of an artificial lightweight geopolymer aggregate (ALGA) using LUSI (Sidoarjo mud) and alkaline activator as source materials. LUSI stands for LU-Lumpur and SI-Sidoarjo, meaning mud from Sidoarjo which erupted near the Banjarpanji-1 exploration well in Sidoarjo, East Java, Indonesia on 27 May 2006. The effect of <span class="hlt">Na</span>OH molarity, LUSI mud/Alkaline activator (LM/AA) ratio, and <span class="hlt">Na</span><span class="hlt">2</span>SiO3/<span class="hlt">Na</span>OH ratio to the ALGA are investigated at a sintering temperature of 950 °C. The results show that the optimum <span class="hlt">Na</span>OH molarity found in this study is 12 M due to the highest strength (lowest AIV value) of 15.79% with lower water absorption and specific gravity. The optimum LUSI mud/Alkaline activator (LM/AA) ratio of 1.7 and the <span class="hlt">Na</span><span class="hlt">2</span>SiO3/<span class="hlt">Na</span>OH ratio of 0.4 gives the highest strength with AIV value of 15.42% with specific gravity of 1.10 g/cm3 and water absorption of 4.7%. The major synthesized crystalline phases were identified as sodalite, quartz and albite. Scanning Electron Microscope (SEM) image showed more complete geopolymer matrix which contributes to highest strength of ALGA produced. PMID:26006238</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JSSCh.248...75A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JSSCh.248...75A"><span>Phase equilibria in the <span class="hlt">NaF-CdO-Na</span>PO3 system at 873 K and crystal structure and physico-chemical characterizations of the new <span class="hlt">Na</span><span class="hlt">2</span>CdPO4F fluorophosphate</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Aboussatar, Mohamed; Mbarek, Aïcha; Naili, Houcine; El-Ghozzi, Malika; Chadeyron, Geneviève; Avignant, Daniel; Zambon, Daniel</p> <p>2017-04-01</p> <p>Isothermal sections of the diagram representing phase relationships in the <span class="hlt">NaF-CdO-Na</span>PO3 system have been investigated by solid state reactions and powder X-ray diffraction. This phase diagram investigation confirms the polymorphism of the <span class="hlt">Na</span>CdPO4 side component and the structure of the ß high temperature polymorph (orthorhombic, space group Pnma and unit cell parameters a=9.3118(<span class="hlt">2</span>), b=7.0459(1), c=5.1849(1) Å has been refined. A new fluorophosphate, <span class="hlt">Na</span><span class="hlt">2</span>CdPO4F, has been discovered and its crystal structure determined and refined from powder X-ray diffraction data. It exhibits a new 3D structure with orthorhombic symmetry, space group Pnma and unit cell parameters a=5.3731(1), b=6.8530(1), c=12.2691(<span class="hlt">2</span>) Å. The structure is closely related to those of the high temperature polymorph of the nacaphite <span class="hlt">Na</span><span class="hlt">2</span>CaPO4F and the fluorosilicate Ca<span class="hlt">2</span><span class="hlt">Na</span>SiO4F but differs essentially in the cationic repartition since the structure is fully ordered with one <span class="hlt">Na</span> site (8d) and one Cd site (4c). Relationships with other <span class="hlt">Na</span><span class="hlt">2</span>MIIPO4F (MII=Mg, Ca, Mn, Fe, Co, Ni) have been examined and the crystal-chemical and topographical analysis of these fluorophosphates is briefly reviewed. IR, Raman, optical and 19F, 23<span class="hlt">Na</span>, 31P MAS NMR characterizations of <span class="hlt">Na</span><span class="hlt">2</span>CdPO4F have been investigated.</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('https://www.ncbi.nlm.nih.gov/pubmed/24449376','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24449376"><span>Effect of salts (<span class="hlt">Na</span>Cl and <span class="hlt">Na</span><span class="hlt">2</span>CO3) on callus and suspension culture of Stevia rebaudiana for Steviol glycoside production.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Gupta, Pratibha; Sharma, Satyawati; Saxena, Sanjay</p> <p>2014-03-01</p> <p>Steviol glycosides are natural non-caloric sweeteners which are extracted from the leaves of Stevia rebaudiana plant. Present study deals the effect of salts (<span class="hlt">Na</span>Cl and <span class="hlt">Na</span><span class="hlt">2</span>CO3) on callus and suspension culture of Stevia plant for steviol glycoside (SGs) production. Yellow-green and compact calli obtained from in vitro raised Stevia leaves sub-cultured on MS medium supplemented with <span class="hlt">2</span>.0 mg l(-1) NAA and different concentrations of <span class="hlt">Na</span>Cl (0.05-0.20%) and <span class="hlt">Na</span><span class="hlt">2</span>CO3 (0.0125-0.10%) for <span class="hlt">2</span> weeks, and incubated at 24 ± 1 °C and 22.4 μmol m(-<span class="hlt">2</span>) s(-1) light intensity provided by white fluorescent tubes for 16 h. Callus and suspension biomass cultured on salts showed less growth as well as browning of medium when compared with control. Quantification of SGs content in callus culture (collected on 15th day) and suspension cultures (collected at 10th and 15th days) treated with and without salts were analyzed by HPLC. It was found that abiotic stress induced by the salts increased the concentration of SGs significantly. In callus, the quantity of SGs got increased from 0.27 (control) to 1.43 and 1.57% with 0.10% <span class="hlt">Na</span>Cl, and 0.025% <span class="hlt">Na</span><span class="hlt">2</span>CO3, respectively. However, in case of suspension culture, the same concentrations of <span class="hlt">Na</span>Cl and <span class="hlt">Na</span><span class="hlt">2</span>CO3 enhanced the SGs content from 1.36 (control) to <span class="hlt">2</span>.61 and 5.14%, respectively, on the 10th day.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3942702','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3942702"><span>Detailed investigation of <span class="hlt">Na</span><span class="hlt">2</span>.24FePO4CO3 as a cathode material for <span class="hlt">Na</span>-ion batteries</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Huang, Weifeng; Zhou, Jing; Li, Biao; Ma, Jin; Tao, Shi; Xia, Dingguo; Chu, Wangsheng; Wu, Ziyu</p> <p>2014-01-01</p> <p><span class="hlt">Na</span>-ion batteries are gaining an increased recognition as the next generation low cost energy storage devices. Here, we present a characterization of <span class="hlt">Na</span>3FePO4CO3 nanoplates as a novel cathode material for sodium ion batteries. First-principles calculations reveal that there are two paths for <span class="hlt">Na</span> ion migration along b and c axis. In-situ and ex-situ Fe K-edge X-ray absorption near edge structure (XANES) point out that in <span class="hlt">Na</span>3FePO4CO3 both Fe<span class="hlt">2</span>+/Fe3+ and Fe3+/Fe4+ redox couples are electrochemically active, suggesting also the existence of a two-electron intercalation reaction. Ex-situ X-ray powder diffraction data demonstrates that the crystalline structure of <span class="hlt">Na</span>3FePO4CO3 remains stable during the charging/discharging process within the range <span class="hlt">2</span>.0–4.55 V. PMID:24595232</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28422389','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28422389"><span>Triosephosphate isomerase tyrosine nitration induced by heme-<span class="hlt">Na</span>NO<span class="hlt">2</span> -H<span class="hlt">2</span> O<span class="hlt">2</span> or peroxynitrite: Effects of different natural phenolic compounds.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Gao, Wanxia; Zhao, Jie; Li, Hailing; Gao, Zhonghong</p> <p>2017-06-01</p> <p>Peroxynitrite and heme peroxidases (or heme)-H <span class="hlt">2</span> O <span class="hlt">2</span> -<span class="hlt">Na</span>NO <span class="hlt">2</span> system are the two common ways to cause protein tyrosine nitration in vitro, but the effects of antioxidants on reducing these two pathways-induced protein nitration and oxidation are controversial. Both nitrating systems can dose-dependently induce triosephosphate isomerase (TIM) nitration, however, heme-H <span class="hlt">2</span> O <span class="hlt">2</span> -<span class="hlt">Na</span>NO <span class="hlt">2</span> was less destructive to protein secondary structures and led to more nitrated tyrosine residue than 3-morpholinosydnonimine hydrochloride (SIN-1, a peroxynitrite donor). Both of desferrioxamine and catechin could inhibit TIM nitration induced by heme-H <span class="hlt">2</span> O <span class="hlt">2</span> -<span class="hlt">Na</span>NO <span class="hlt">2</span> and SIN-1 and protein oxidation induced by SIN-1, but promoted heme-H <span class="hlt">2</span> O <span class="hlt">2</span> -<span class="hlt">Na</span>NO <span class="hlt">2</span> -induced protein oxidation. Moreover, the antagonism of natural phenolic compounds on SIN-1-induced tyrosine nitration was consistent with their radical scavenging ability, but no similar consensus was found in heme-H <span class="hlt">2</span> O <span class="hlt">2</span> -<span class="hlt">Na</span>NO <span class="hlt">2</span> -induced nitration. Our results indicated that peroxynitrite and heme-H <span class="hlt">2</span> O <span class="hlt">2</span> -<span class="hlt">Na</span>NO <span class="hlt">2</span> -induced protein nitration was different, and the later one could be a better model for anti-nitration compounds screening. © 2017 Wiley Periodicals, Inc.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29671750','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29671750"><span>Thermoelectric properties of layered <span class="hlt">Na</span>SbSe<span class="hlt">2</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Putatunda, Aditya; Xing, Guangzong; Sun, Jifeng; Li, Yuwei; Singh, David J</p> <p>2018-06-06</p> <p>We investigate ordered monoclinic <span class="hlt">Na</span>SbSe <span class="hlt">2</span> as a thermoelectric using first principles calculations. We find that from an electronic point of view, ordered and oriented n-type <span class="hlt">Na</span>SbSe <span class="hlt">2</span> is comparable to the best known thermoelectric materials. This phase has a sufficiently large band gap for thermoelectric and solar absorber applications in contrast to the disordered phase which has a much narrower gap. The electronic structure shows anisotropic, non-parabolic bands. The results show a high Seebeck coefficient in addition to direction dependent high conductivity. The electronic structure quantified by an electron fitness function is very favorable, especially in the n-type case.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JPCM...30v5501P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JPCM...30v5501P"><span>Thermoelectric properties of layered <span class="hlt">Na</span>SbSe<span class="hlt">2</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Putatunda, Aditya; Xing, Guangzong; Sun, Jifeng; Li, Yuwei; Singh, David J.</p> <p>2018-06-01</p> <p>We investigate ordered monoclinic <span class="hlt">Na</span>SbSe<span class="hlt">2</span> as a thermoelectric using first principles calculations. We find that from an electronic point of view, ordered and oriented n-type <span class="hlt">Na</span>SbSe<span class="hlt">2</span> is comparable to the best known thermoelectric materials. This phase has a sufficiently large band gap for thermoelectric and solar absorber applications in contrast to the disordered phase which has a much narrower gap. The electronic structure shows anisotropic, non-parabolic bands. The results show a high Seebeck coefficient in addition to direction dependent high conductivity. The electronic structure quantified by an electron fitness function is very favorable, especially in the n-type case.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JPCS..114..201Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JPCS..114..201Z"><span><span class="hlt">Na</span>3Tb(PO4)<span class="hlt">2</span>: Synthesis, crystal structure and greenish emitting properties</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zhao, Dan; Ma, Zhao; Liu, Bao-Zhong; Zhang, Rui-Juan; Wu, Zhi-Qiang; Wang, Jian; Duan, Pei-Gao</p> <p>2018-03-01</p> <p>A anhydrous orthoborate <span class="hlt">Na</span>3Tb(PO4)<span class="hlt">2</span> has been prepared and its crystal structure was determined by X-Ray diffraction of a non-merohedral twinned single crystal. The results show that the compound crystallizes in monoclinic space group C<span class="hlt">2</span>/c and the structure features a 3D framework containing PO4, <span class="hlt">Na</span>O6, <span class="hlt">Na</span>O7, <span class="hlt">Na</span>O8 and TbO8 polyhedra. Under near-UV excitation (370 nm), <span class="hlt">Na</span>3Tb(PO4)<span class="hlt">2</span> shows intense characteristic emission bands of Tb3+ (490 nm, 543 nm, 585 nm and 620 nm) with the CIE coordinate of (0.3062, 0.5901), corresponding to greenish color. The excitation spectrum covers a wide range from 340 nm to 390 nm, which indicates that phosphor <span class="hlt">Na</span>3Tb(PO4)<span class="hlt">2</span> can be efficiently activated by near-UV LED ship.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27700032','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27700032"><span>Sodium Ion Diffusion in Nasicon (<span class="hlt">Na</span>3Zr<span class="hlt">2</span>Si<span class="hlt">2</span>PO12) Solid Electrolytes: Effects of Excess Sodium.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Park, Heetaek; Jung, Keeyoung; Nezafati, Marjan; Kim, Chang-Soo; Kang, Byoungwoo</p> <p>2016-10-04</p> <p>The <span class="hlt">Na</span> superionic conductor (aka Nasicon, <span class="hlt">Na</span> 1+x Zr <span class="hlt">2</span> Si x P 3-x O 12 , where 0 ≤ x ≤ 3) is one of the promising solid electrolyte materials used in advanced molten <span class="hlt">Na</span>-based secondary batteries that typically operate at high temperature (over ∼270 °C). Nasicon provides a 3D diffusion network allowing the transport of the active <span class="hlt">Na</span>-ion species (i.e., ionic conductor) while blocking the conduction of electrons (i.e., electronic insulator) between the anode and cathode compartments of cells. In this work, the standard Nasicon (<span class="hlt">Na</span> 3 Zr <span class="hlt">2</span> Si <span class="hlt">2</span> PO 12 , bare sample) and 10 at% <span class="hlt">Na</span>-excess Nasicon (<span class="hlt">Na</span> 3.3 Zr <span class="hlt">2</span> Si <span class="hlt">2</span> PO 12 , <span class="hlt">Na</span>-excess sample) solid electrolytes were synthesized using a solid-state sintering technique to elucidate the <span class="hlt">Na</span> diffusion mechanism (i.e., grain diffusion or grain boundary diffusion) and the impacts of adding excess <span class="hlt">Na</span> at relatively low and high temperatures. The structural, thermal, and ionic transport characterizations were conducted using various experimental tools including X-ray diffraction (XRD), differential scanning calorimetry (DSC), scanning electron microscopy (SEM), and electrochemical impedance spectroscopy (EIS). In addition, an ab initio atomistic modeling study was carried out to computationally examine the detailed microstructures of Nasicon materials, as well as to support the experimental observations. Through this combination work comprising experimental and computational investigations, we show that the predominant mechanisms of <span class="hlt">Na</span>-ion transport in the Nasicon structure are the grain boundary and the grain diffusion at low and high temperatures, respectively. Also, it was found that adding 10 at% excess <span class="hlt">Na</span> could give rise to a substantial increase in the total conductivity (e.g., ∼1.<span class="hlt">2</span> × 10 -1 S/cm at 300 °C) of Nasicon electrolytes resulting from the enlargement of the bottleneck areas in the <span class="hlt">Na</span> diffusion channels of polycrystalline grains.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22658241-phase-equilibria-naf-cdo-napo-sub-system-crystal-structure-physico-chemical-characterizations-new-na-sub-cdpo-sub-fluorophosphate','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22658241-phase-equilibria-naf-cdo-napo-sub-system-crystal-structure-physico-chemical-characterizations-new-na-sub-cdpo-sub-fluorophosphate"><span>Phase equilibria in the <span class="hlt">NaF-CdO-Na</span>PO{sub 3} system at 873 K and crystal structure and physico-chemical characterizations of the new <span class="hlt">Na</span>{sub <span class="hlt">2</span>}CdPO{sub 4}F fluorophosphate</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>Aboussatar, Mohamed; Laboratoire de Physico-Chimie de l’État Solide, Faculté des Sciences de Sfax, Université de Sfax, BP 1171, 3000 Sfax; Mbarek, Aïcha</p> <p></p> <p>Isothermal sections of the diagram representing phase relationships in the <span class="hlt">NaF-CdO-Na</span>PO{sub 3} system have been investigated by solid state reactions and powder X-ray diffraction. This phase diagram investigation confirms the polymorphism of the <span class="hlt">Na</span>CdPO{sub 4} side component and the structure of the ß high temperature polymorph (orthorhombic, space group Pnma and unit cell parameters a=9.3118(<span class="hlt">2</span>), b=7.0459(1), c=5.1849(1) Å has been refined. A new fluorophosphate, <span class="hlt">Na</span>{sub <span class="hlt">2</span>}CdPO{sub 4}F, has been discovered and its crystal structure determined and refined from powder X-ray diffraction data. It exhibits a new 3D structure with orthorhombic symmetry, space group Pnma and unit cell parameters a=5.3731(1), b=6.8530(1),more » c=12.2691(<span class="hlt">2</span>) Å. The structure is closely related to those of the high temperature polymorph of the nacaphite <span class="hlt">Na</span>{sub <span class="hlt">2</span>}CaPO{sub 4}F and the fluorosilicate Ca{sub <span class="hlt">2</span>}<span class="hlt">Na</span>SiO{sub 4}F but differs essentially in the cationic repartition since the structure is fully ordered with one <span class="hlt">Na</span> site (8d) and one Cd site (4c). Relationships with other <span class="hlt">Na</span>{sub <span class="hlt">2</span>}M{sup II}PO{sub 4}F (M{sup II}=Mg, Ca, Mn, Fe, Co, Ni) have been examined and the crystal-chemical and topographical analysis of these fluorophosphates is briefly reviewed. IR, Raman, optical and {sup 19}F, {sup 23}<span class="hlt">Na</span>, {sup 31}P MAS NMR characterizations of <span class="hlt">Na</span>{sub <span class="hlt">2</span>}CdPO{sub 4}F have been investigated. - Graphical abstract: The structure of the compound <span class="hlt">Na</span>{sub <span class="hlt">2</span>}CdPO{sub 4}F, discovered during the study of the phase relationships in the <span class="hlt">NaF-CdO-Na</span>PO{sub 3} system, has been determined and compared with other <span class="hlt">Na</span>{sub <span class="hlt">2</span>}M{sup II}PO{sub 4}F fluorophosphates. - Highlights: • XRD analysis of the isothermal section of the <span class="hlt">NaF-CdO-Na</span>PO{sub 3} system at 923 K. • Rietveld refinement of the high temperature polymorph β-<span class="hlt">Na</span>CdPO{sub 4}. • Crystal structure of the new <span class="hlt">Na</span>{sub <span class="hlt">2</span>}CdPO{sub 4}F fluorophosphate determined from powder XRD. • Crystal structure - composition relationships of <span class="hlt">Na</span>{sub <span class="hlt">2</span>}M{sup II}PO{sub 4}F</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018RJPCA..92.1213L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018RJPCA..92.1213L"><span>Thermodynamic Study of Solid-Liquid Equilibrium in <span class="hlt">NaCl-Na</span>Br-H<span class="hlt">2</span>O System at 288.15 K</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Li, Dan; Meng, Ling-zong; Deng, Tian-long; Guo, Ya-fei; Fu, Qing-Tao</p> <p>2018-06-01</p> <p>The solubility data, composition of the solid solution and refractive indices of the <span class="hlt">NaCl-Na</span>Br-H<span class="hlt">2</span>O system at 288.15 K were studied with the isothermal equilibrium dissolution method. The solubility diagram and refractive index diagram of this system were plotted at 288.15 K. The solubility diagram consists of two crystallization zones for solid solution <span class="hlt">Na</span>(Cl,Br) · <span class="hlt">2</span>H<span class="hlt">2</span>O and <span class="hlt">Na</span>(Cl,Br), one invariant points cosaturated with two solid solution and two univariant solubility isothermal curves. On the basis of Pitzer and Harvie-Weare (HW) chemical models, the composition equations and solubility equilibrium constant equations of the solid solutions at 288.15 K were acquired using the solubility data, the composition of solid solutions, and binary Pitzer parameters. The solubilities calculated using the new method combining the equations are in good agreement with the experimental data.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27337045','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27337045"><span>Identification of a 3rd <span class="hlt">Na</span>+ Binding Site of the Glycine Transporter, GlyT<span class="hlt">2</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Subramanian, Nandhitha; Scopelitti, Amanda J; Carland, Jane E; Ryan, Renae M; O'Mara, Megan L; Vandenberg, Robert J</p> <p>2016-01-01</p> <p>The <span class="hlt">Na</span>+/Cl- dependent glycine transporters GlyT1 and GlyT<span class="hlt">2</span> regulate synaptic glycine concentrations. Glycine transport by GlyT<span class="hlt">2</span> is coupled to the co-transport of three <span class="hlt">Na</span>+ ions, whereas transport by GlyT1 is coupled to the co-transport of only two <span class="hlt">Na</span>+ ions. These differences in ion-flux coupling determine their respective concentrating capacities and have a direct bearing on their functional roles in synaptic transmission. The crystal structures of the closely related bacterial <span class="hlt">Na</span>+-dependent leucine transporter, LeuTAa, and the Drosophila dopamine transporter, dDAT, have allowed prediction of two <span class="hlt">Na</span>+ binding sites in GlyT<span class="hlt">2</span>, but the physical location of the third <span class="hlt">Na</span>+ site in GlyT<span class="hlt">2</span> is unknown. A bacterial betaine transporter, BetP, has also been crystallized and shows structural similarity to LeuTAa. Although betaine transport by BetP is coupled to the co-transport of two <span class="hlt">Na</span>+ ions, the first <span class="hlt">Na</span>+ site is not conserved between BetP and LeuTAa, the so called <span class="hlt">Na</span>1' site. We hypothesized that the third <span class="hlt">Na</span>+ binding site (<span class="hlt">Na</span>3 site) of GlyT<span class="hlt">2</span> corresponds to the BetP <span class="hlt">Na</span>1' binding site. To identify the <span class="hlt">Na</span>3 binding site of GlyT<span class="hlt">2</span>, we performed molecular dynamics (MD) simulations. Surprisingly, a <span class="hlt">Na</span>+ placed at the location consistent with the <span class="hlt">Na</span>1' site of BetP spontaneously dissociated from its initial location and bound instead to a novel <span class="hlt">Na</span>3 site. Using a combination of MD simulations of a comparative model of GlyT<span class="hlt">2</span> together with an analysis of the functional properties of wild type and mutant GlyTs we have identified an electrostatically favorable novel third <span class="hlt">Na</span>+ binding site in GlyT<span class="hlt">2</span> formed by Trp263 and Met276 in TM3, Ala481 in TM6 and Glu648 in TM10.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4232491','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4232491"><span>Analysis of the <span class="hlt">Na</span>+/Ca<span class="hlt">2</span>+ Exchanger Gene Family within the Phylum Nematoda</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>He, Chao; O'Halloran, Damien M.</p> <p>2014-01-01</p> <p><span class="hlt">Na</span>+/Ca<span class="hlt">2</span>+ exchangers are low affinity, high capacity transporters that rapidly transport calcium at the plasma membrane, mitochondrion, endoplasmic (and sarcoplasmic) reticulum, and the nucleus. <span class="hlt">Na</span>+/Ca<span class="hlt">2</span>+ exchangers are widely expressed in diverse cell types where they contribute homeostatic balance to calcium levels. In animals, <span class="hlt">Na</span>+/Ca<span class="hlt">2</span>+ exchangers are divided into three groups based upon stoichiometry: <span class="hlt">Na</span>+/Ca<span class="hlt">2</span>+ exchangers (NCX), <span class="hlt">Na</span>+/Ca<span class="hlt">2</span>+/K+ exchangers (NCKX), and Ca<span class="hlt">2</span>+/Cation exchangers (CCX). In mammals there are three NCX genes, five NCKX genes and one CCX (NCLX) gene. The genome of the nematode Caenorhabditis elegans contains ten <span class="hlt">Na</span>+/Ca<span class="hlt">2</span>+ exchanger genes: three NCX; five CCX; and two NCKX genes. Here we set out to characterize structural and taxonomic specializations within the family of <span class="hlt">Na</span>+/Ca<span class="hlt">2</span>+ exchangers across the phylum Nematoda. In this analysis we identify <span class="hlt">Na</span>+/Ca<span class="hlt">2</span>+ exchanger genes from twelve species of nematodes and reconstruct their phylogenetic and evolutionary relationships. The most notable feature of the resulting phylogenies was the heterogeneous evolution observed within exchanger subtypes. Specifically, in the case of the CCX exchangers we did not detect members of this class in three Clade III nematodes. Within the Caenorhabditis and Pristionchus lineages we identify between three and five CCX representatives, whereas in other Clade V and also Clade IV nematode taxa we only observed a single CCX gene in each species, and in the Clade III nematode taxa that we sampled we identify NCX and NCKX encoding genes but no evidence of CCX representatives using our mining approach. We also provided re-annotation for predicted CCX gene structures from Heterorhabditis bacteriophora and Caenorhabditis japonica by RT-PCR and sequencing. Together, these findings reveal a complex picture of <span class="hlt">Na</span>+/Ca<span class="hlt">2</span>+ transporters in nematodes that suggest an incongruent evolutionary history of proteins that provide central control of calcium dynamics. PMID:25397810</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015ApSS..347..401G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015ApSS..347..401G"><span>Electrochemical preparation of carbon films with a Mo<span class="hlt">2</span>C interlayer in LiCl-<span class="hlt">NaCl-Na</span><span class="hlt">2</span>CO3 melts</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ge, Jianbang; Wang, Shuai; Zhang, Feng; Zhang, Long; Jiao, Handong; Zhu, Hongmin; Jiao, Shuqiang</p> <p>2015-08-01</p> <p>The electrodeposition of carbon films with a Mo<span class="hlt">2</span>C interlayer was investigated in LiCl-<span class="hlt">NaCl-Na</span><span class="hlt">2</span>CO3 melts at 900 °C. Cyclic voltammetry was applied to study the electrochemical reaction mechanism on Mo and Pt electrodes, indicating that, two reduction reactions including carbon deposition and carbon monoxide evolution, may take place on the two electrodes simultaneously during the cathodic sweep. Carbon films with a continuous Mo<span class="hlt">2</span>C interlayer were prepared by constant voltage electrolysis, showing a good adhesion between Mo substrate and carbon films. The carbon films with a Mo<span class="hlt">2</span>C interlayer were characterized using X-ray diffraction measurement, Raman spectroscopy, scanning electron microscopy and transmission electron microscopy. The results reveal that carbon materials deposited on the electrodes are mainly composed of graphite and carbon diffusion in Mo (or Mo<span class="hlt">2</span>C) leads to the formation and growth of Mo<span class="hlt">2</span>C interlayer.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29159314','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29159314"><span>Vasoconstriction triggered by hydrogen sulfide: Evidence for <span class="hlt">Na</span>+,K+,<span class="hlt">2</span>Cl-cotransport and L-type Ca<span class="hlt">2</span>+ channel-mediated pathway.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Orlov, Sergei N; Gusakova, Svetlana V; Smaglii, Liudmila V; Koltsova, Svetlana V; Sidorenko, Svetalana V</p> <p>2017-12-01</p> <p>This study examined the dose-dependent actions of hydrogen sulfide donor sodium hydrosulphide (<span class="hlt">Na</span>HS) on isometric contractions and ion transport in rat aorta smooth muscle cells (SMC). Isometric contraction was measured in ring aortas segments from male Wistar rats. Activity of <span class="hlt">Na</span> + /K + -pump and <span class="hlt">Na</span> + ,K + ,<span class="hlt">2</span>Cl - cotransport was measured in cultured endothelial and smooth muscle cells from the rat aorta as ouabain-sensitive and ouabain-resistant, bumetanide-sensitive components of the 86 Rb influx, respectively. <span class="hlt">Na</span>HS exhibited the bimodal action on contractions triggered by modest depolarization ([K + ] o =30 mM). At 10 -4 M, <span class="hlt">Na</span>HS augmented contractions of intact and endothelium-denuded strips by ~ 15% and 25%, respectively, whereas at concentration of 10 -3  M it decreased contractile responses by more than two-fold. Contractions evoked by 10 -4  M <span class="hlt">Na</span>HS were completely abolished by bumetanide, a potent inhibitor of <span class="hlt">Na</span> + ,K + ,<span class="hlt">2</span>Cl - cotransport, whereas the inhibition seen at 10 -3  M <span class="hlt">Na</span>HS was suppressed in the presence of K + channel blocker TEA. In cultured SMC, 5×10 -5  M <span class="hlt">Na</span>HS increased <span class="hlt">Na</span> + ,K + ,<span class="hlt">2</span>Cl - - cotransport without any effect on the activity of this carrier in endothelial cells. In depolarized SMC, 45 Ca influx was enhanced in the presence of 10 -4  M <span class="hlt">Na</span>HS and suppressed under elevation of [<span class="hlt">Na</span>HS] up to 10 -3  M. 45 Ca influx triggered by 10 -4  M <span class="hlt">Na</span>HS was abolished by bumetanide and L-type Ca <span class="hlt">2</span>+ channel blocker nicardipine. Our results strongly suggest that contractions of rat aortic rings triggered by low doses of <span class="hlt">Na</span>HS are mediated by activation of <span class="hlt">Na</span> + ,K + ,<span class="hlt">2</span>Cl - cotransport and Ca <span class="hlt">2</span>+ influx via L-type channels.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/16929642','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/16929642"><span>Organic pollution and salt intrusion in <span class="hlt">Cai</span> Nuoc District, Ca Mau Province, Vietnam.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Tho, Nguyen; Vromant, Nico; Hung, Nguyen Thanh; Hens, Luc</p> <p>2006-07-01</p> <p>In Ca Mau, Vietnam, farmers converted from rice to shrimp farming, while ignoring the degradation of the aquatic environment. We assessed the seasonal variations in organic pollution of the surface water and salt intrusion in one district and assessed the difference in chemical characteristics of the surface water of shrimp ponds and canals. Several variables reflecting salinity and organic pollution were measured in the wet and dry season. The results show that in the dry season salinity increased to 37.36-42.73 g l(-1) and COD and suspended solids increased to a maximum of 268.7 mg l(-1) and 1312.0 mg l(-1), respectively. In the wet season salinity values of 8.16 to 10.60 g l(-1) were recorded, indicating that salinity could no longer be washed out completely in this season. It is concluded that salinity and suspended solids in the aquatic environment in the <span class="hlt">Cai</span> Nuoc district are increased by shrimp monoculture, whereas organic pollution is contributed by human population pressure.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27208740','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27208740"><span>Enhanced enzymatic saccharification of sugarcane bagasse pretreated by combining O<span class="hlt">2</span> and <span class="hlt">Na</span>OH.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Bi, Shuaizhu; Peng, Lincai; Chen, Keli; Zhu, Zhengliang</p> <p>2016-08-01</p> <p>Sugarcane bagasse pretreated by combining O<span class="hlt">2</span> and <span class="hlt">Na</span>OH with different variables was conducted to improve its enzymatic digestibility and sugar recovery, and the results were compared with sole <span class="hlt">Na</span>OH pretreatment. Lignin removal for O<span class="hlt">2</span>-<span class="hlt">Na</span>OH pretreatment was around 10% higher than that for sole <span class="hlt">Na</span>OH pretreatment under the same conditions, and O<span class="hlt">2</span>-<span class="hlt">Na</span>OH pretreatment resulted in higher glucan recovery in the solid remain. Subsequently, O<span class="hlt">2</span>-<span class="hlt">Na</span>OH pretreated sugarcane bagasse presented more efficient enzymatic digestibility than sole <span class="hlt">Na</span>OH pretreatment. Under the moderate pretreatment conditions of combining 1% <span class="hlt">Na</span>OH and 0.5MPa O<span class="hlt">2</span> at 80°C for 120min, a high glucan conversion of 95% was achieved after 48h enzymatic hydrolysis. Coupled with the operations of pretreatment and enzymatic hydrolysis, an admirable total sugar recovery of 89% (glucose recovery of 93% and xylose recovery of 84%) was obtained. The susceptibility of the substrates to enzymatic digestibility was explained by their physical and chemical characteristics. Copyright © 2016 Elsevier Ltd. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22494920-robust-high-pressure-stability-negative-thermal-expansion-sodium-rich-antiperovskites-na-sub-obr-na-sub-oi-sub','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22494920-robust-high-pressure-stability-negative-thermal-expansion-sodium-rich-antiperovskites-na-sub-obr-na-sub-oi-sub"><span>Robust high pressure stability and negative thermal expansion in sodium-rich antiperovskites <span class="hlt">Na</span>{sub 3}OBr and <span class="hlt">Na</span>{sub 4}OI{sub <span class="hlt">2</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>Wang, Yonggang, E-mail: yyggwang@gmail.com, E-mail: yangwg@hpstar.ac.cn, E-mail: yusheng.zhao@unlv.edu; Institute of Nanostructured Functional Materials, Huanghe Science and Technology College, Zhengzhou, Henan 450006; High Pressure Synergetic Consortium</p> <p>2016-01-14</p> <p>The structure stability under high pressure and thermal expansion behavior of <span class="hlt">Na</span>{sub 3}OBr and <span class="hlt">Na</span>{sub 4}OI{sub <span class="hlt">2</span>}, two prototypes of alkali-metal-rich antiperovskites, were investigated by in situ synchrotron X-ray diffraction techniques under high pressure and low temperature. Both are soft materials with bulk modulus of 58.6 GPa and 52.0 GPa for <span class="hlt">Na</span>{sub 3}OBr and <span class="hlt">Na</span>{sub 4}OI{sub <span class="hlt">2</span>}, respectively. The cubic <span class="hlt">Na</span>{sub 3}OBr structure and tetragonal <span class="hlt">Na</span>{sub 4}OI{sub <span class="hlt">2</span>} with intergrowth K{sub <span class="hlt">2</span>}NiF{sub 4} structure are stable under high pressure up to 23 GPa. Although being a characteristic layered structure, <span class="hlt">Na</span>{sub 4}OI{sub <span class="hlt">2</span>} exhibits nearly isotropic compressibility. Negative thermal expansion was observed at lowmore » temperature range (20–80 K) in both transition-metal-free antiperovskites for the first time. The robust high pressure structure stability was examined and confirmed by first-principles calculations among various possible polymorphisms qualitatively. The results provide in-depth understanding of the negative thermal expansion and robust crystal structure stability of these antiperovskite systems and their potential applications.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014APS..MARD24009A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014APS..MARD24009A"><span><span class="hlt">Na</span>0.44MnO<span class="hlt">2</span> nanorods as a cathode material for <span class="hlt">Na</span>-ion batteries</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Avci, Sevda; Oz, Erdinc; Demirel, Serkan; Altin, Emine; Altin, Serdar; Bayri, Ali; Yakinci, Eyyuphan</p> <p>2014-03-01</p> <p>Lithium-ion batteries have dominated the rechargeable battery market because of their high energy and power capability. On the other hand, sodium is one of the more abundant elements on Earth unlike Li. Moreover, <span class="hlt">Na</span> has similar chemical properties to Li, indicating that <span class="hlt">Na</span>-ion batteries can be an alternative to Li counterparts. With that respect, we have synthesized <span class="hlt">Na</span>0.44MnO<span class="hlt">2</span> nanorods as cathode materials for <span class="hlt">Na</span>-ion batteries. We have investigated the effects of structural, electrical, and magnetic properties on battery performance. We report the synthesis conditions and growth mechanism of the nanorods. The structure and the morphology of the materials were investigated by X-ray diffraction (XRD), scanning electron microscopy (SEM), energy-dispersive X-ray (EDX), and atomic force microscopy (AFM) techniques. Temperature dependent structural changes were determined via in situ X-ray diffraction and TG-DTA measurements showing structural changes above room temperature. This work is funded by The Scientific and Technological Research Council of Turkey with Grant No:112M487.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24255867','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24255867"><span>Plant growth-promoting activities of Streptomyces spp. in sorghum and rice.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Gopalakrishnan, Subramaniam; Srinivas, Vadlamudi; Sree Vidya, Meesala; Rathore, Abhishek</p> <p>2013-01-01</p> <p>Five strains of Streptomyces (<span class="hlt">CAI</span>-24, <span class="hlt">CAI</span>-121, <span class="hlt">CAI</span>-127, KAI-32 and KAI-90) were earlier reported by us as biological control agents against Fusarium wilt of chickpea caused by Fusarium oxysporum f. sp. ciceri (FOC). In the present study, the Streptomyces were characterized for enzymatic activities, physiological traits and further evaluated in greenhouse and field for their plant growth promotion (PGP) of sorghum and rice. All the Streptomyces produced lipase, β-1-3-glucanase and chitinase (except <span class="hlt">CAI</span>-121 and <span class="hlt">CAI</span>-127), grew in <span class="hlt">Na</span>Cl concentrations of up to 6%, at pH values between 5 and 13 and temperatures between 20 and 40°C and were highly sensitive to Thiram, Benlate, Captan, Benomyl and Radonil at field application level. When the Streptomyces were evaluated in the greenhouse on sorghum all the isolates significantly enhanced all the agronomic traits over the control. In the field, on rice, the Streptomyces significantly enhanced stover yield (up to 25%; except <span class="hlt">CAI</span>-24), grain yield (up to 10%), total dry matter (up to 18%; except <span class="hlt">CAI</span>-24) and root length, volume and dry weight (up to 15%, 36% and 55%, respectively, except <span class="hlt">CAI</span>-24) over the control. In the rhizosphere soil, the Streptomyces significantly enhanced microbial biomass carbon (except <span class="hlt">CAI</span>-24), nitrogen, dehydrogenase (except <span class="hlt">CAI</span>-24), total N, available P and organic carbon (up to 41%, 52%, 75%, 122%, 53% and 13%, respectively) over the control. This study demonstrates that the selected Streptomyces which were antagonistic to FOC also have PGP properties.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1251445-layered-rock-salt-transformation-desodiated-naxcro2','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1251445-layered-rock-salt-transformation-desodiated-naxcro2"><span>Layered-to-Rock-Salt Transformation in Desodiated <span class="hlt">Na</span> xCrO <span class="hlt">2</span> ( x 0.4)</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Bo, Shou-Hang; Li, Xin; Toumar, Alexandra J.; ...</p> <p>2016-02-01</p> <p>O3 layered sodium transition metal oxides (i.e., <span class="hlt">Na</span>MO <span class="hlt">2</span>, M = Ti, V, Cr, Mn, Fe, Co, Ni) are a promising class of cathode materials for <span class="hlt">Na</span>-ion battery applications. These materials, however, all suffer from severe capacity decay when the extraction of <span class="hlt">Na</span> exceeds certain capacity limits. Understanding the causes of this capacity decay is critical to unlocking the potential of these materials for battery applications. In this work, we investigate the structural origins of capacity decay for one of the compounds in this class, <span class="hlt">Na</span>CrO <span class="hlt">2</span>. The (de)sodiation processes of <span class="hlt">Na</span>CrO <span class="hlt">2</span> were studied both in situ and exmore » situ through X-ray and electron diffraction measurements. We demonstrate that <span class="hlt">Na</span> xCrO <span class="hlt">2</span> (0 < x < 1) remains in the layered structural framework without Cr migration up to a composition of <span class="hlt">Na</span> 0.4CrO <span class="hlt">2</span>. Further removal of <span class="hlt">Na</span> beyond this composition triggers a layered-to-rock-salt transformation, which converts P'3-<span class="hlt">Na</span> 0.4CrO <span class="hlt">2</span> into the rock-salt CrO <span class="hlt">2</span> phase. This structural transformation proceeds via the formation of an intermediate O3 <span class="hlt">Na</span> δCrO <span class="hlt">2</span> phase that contains Cr in both <span class="hlt">Na</span> and Cr slabs and shares very similar lattice dimensions with those of rock-salt CrO <span class="hlt">2</span>. It is intriguing to note that intercalation of alkaline ions (i.e., <span class="hlt">Na</span> + and Li + ) into the rock-salt CrO <span class="hlt">2</span> and O3 <span class="hlt">Na</span> δCrO <span class="hlt">2</span> structures is actually possible, albeit in a limited amount (~0.<span class="hlt">2</span> per formula unit). When these results were analyzed under the context of electrochemistry data, it was apparent that preventing the layered-to-rock-salt transformation is crucial to improve the cyclability of <span class="hlt">Na</span>CrO <span class="hlt">2</span>. Possible strategies for mitigating this detrimental phase transition are proposed.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014APS..MARD24010O','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014APS..MARD24010O"><span>First-principles study of the electronic properties and discharge profile of Ag<span class="hlt">Na</span>(VO<span class="hlt">2</span>F<span class="hlt">2)2</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Onoue, Masatoshi; Trimarchi, Giancarlo; Freeman, Arthur J.</p> <p>2014-03-01</p> <p>Implantable cardiac defibrillators (ICDs) require batteries with high capacities and high discharge rates to ensure the optimal operation of the device over several years. Ag<span class="hlt">2</span>V4O11 has been a cathode material of choice for the ICDs owing to its high capacity and fast rate of electronic discharge. To reduce ICD size and improve ICD performance, a new cathode material would need to display a higher volumetric capacity and redox potential. Recently, the new cathode compound Ag<span class="hlt">Na</span>(VO<span class="hlt">2</span>F<span class="hlt">2)2</span> (SSVOF) was synthesized and displayed favorable voltage for sodium-ion batteries. However, the discharge reaction has been unclear. In this presentation, we study the discharge reaction of SSVOF through DFT calculations. All calculations are performed within the PAW method using the GGA and GGA + U functionals. Among several possible reactions, we focus on the reaction Ag X + A --> AX + Ag, where X is <span class="hlt">Na</span>(VO<span class="hlt">2</span>F<span class="hlt">2)2</span> and A is Li or <span class="hlt">Na</span>. In this reaction, the discharge occurs by replacing Ag with A. The calculated discharge potential for Li is 3.3 V in GGA and <span class="hlt">2</span>.9 V in GGA + U and that for <span class="hlt">Na</span> is 3.1 V in GGA and <span class="hlt">2</span>.8 V in GGA + U . These values are consistent with the experimental ones. Supported by the DOE ER46536 Program.</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/2007GeCoA..71.3557C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2007GeCoA..71.3557C"><span>An isopiestic study of aqueous <span class="hlt">Na</span>Br and KBr at 50 °C: Chemical equilibrium model of solution behavior and solubility in the <span class="hlt">Na</span>Br-H <span class="hlt">2</span>O, KBr-H <span class="hlt">2</span>O and <span class="hlt">Na</span>-K-Br-H <span class="hlt">2</span>O systems to high concentration and temperature</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Christov, Christomir</p> <p>2007-07-01</p> <p>The isopiestic method has been used to determine the osmotic coefficients of the binary solutions <span class="hlt">Na</span>Br-H <span class="hlt">2</span>O (from 0.745 to 5.953 mol kg -1) and KBr-H <span class="hlt">2</span>O (from 0.741 to 5.683 mol kg -1) at the temperature t = 50 °C. Sodium chloride solutions have been used as isopiestic reference standards. The isopiestic results obtained have been combined with all other experimental thermodynamic quantities available in literature (osmotic coefficients, water activities, bromide mineral's solubilities) to construct a chemical model that calculates solute and solvent activities and solid-liquid equilibria in the <span class="hlt">Na</span>Br-H <span class="hlt">2</span>O, KBr-H <span class="hlt">2</span>O and <span class="hlt">Na</span>-K-Br-H <span class="hlt">2</span>O systems from dilute to high solution concentration within the 0-300 °C temperature range. The Harvie and Weare [Harvie C., and Weare J. (1980) The prediction of mineral solubilities in naturalwaters: the <span class="hlt">Na</span>-K-Mg-Ca-Cl-SO 4-H <span class="hlt">2</span>O system from zero to high concentration at 25 °C. Geochim. Cosmochim. Acta44, 981-997] solubility modeling approach, incorporating their implementation of the concentration-dependent specific interaction equations of Pitzer [Pitzer K. (1973) Thermodynamics of electrolytes. I. Theoretical basis and general equations. J. Phys. Chem.77, 268-277] is employed. The model for binary systems is validated by comparing activity coefficient predictions with those given in literature, and not used in the parameterization process. Limitations of the mixed solutions model due to data insufficiencies are discussed. This model expands the variable temperature sodium-potassium model of Greenberg and Moller [Greenberg J., and Moller N. (1989) The prediction of mineral solubilities in natural waters: a chemical equilibrium model for the <span class="hlt">Na</span>-K-Ca-Cl-SO 4-H <span class="hlt">2</span>O system to high concentration from 0 to 250 °C. Geochim. Cosmochim. Acta53, 2503-2518] by evaluating Br - pure electrolyte and mixing solution parameters and the chemical potentials of three bromide solid phases: <span class="hlt">Na</span>Br-<span class="hlt">2</span>H <span class="hlt">2</span>O (cr), <span class="hlt">Na</span>Br (cr) and KBr (cr).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/5956684-role-na-sub-anoxygenic-photosynthesis-sub-production-cyanobacterium-nostoc-muscorum','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/5956684-role-na-sub-anoxygenic-photosynthesis-sub-production-cyanobacterium-nostoc-muscorum"><span>Role of <span class="hlt">Na</span>/sub <span class="hlt">2</span>/S in anoxygenic photosynthesis and H/sub <span class="hlt">2</span>/ production in the cyanobacterium Nostoc Muscorum</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>Fry, I.; Robinson, A.E.; Spath, S.</p> <p>1984-09-28</p> <p><span class="hlt">Na</span>/sub <span class="hlt">2</span>/S is known to support anoxygenic photosynthesis in some strains of cyanobacteria and to stimulate H/sub <span class="hlt">2</span>/ production in N/sub <span class="hlt">2</span>/ fixing filaments of Nostoc muscorum. We have shown electron transfer between <span class="hlt">Na</span>/sub <span class="hlt">2</span>/S and Photosystem I to be dependent on cytochrome b/sub 559/ which was detected only in vegetative cells. An electron mediator was required to support <span class="hlt">Na</span>/sub <span class="hlt">2</span>/S driven nitrogenase activity in isolated heterocysts. <span class="hlt">Na</span>/sub <span class="hlt">2</span>/S was also found to deplete the ATP pool, probably by inhibiting electron transfer from Photosystem I. 14 references, 4 figures.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23340053','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23340053"><span>Magnetic zeolite <span class="hlt">Na</span>A: synthesis, characterization based on metakaolin and its application for the removal of Cu<span class="hlt">2</span>+, Pb<span class="hlt">2</span>+.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Liu, Haibo; Peng, Shuchuan; Shu, Lin; Chen, Tianhu; Bao, Teng; Frost, Ray L</p> <p>2013-06-01</p> <p>The optimum parameters for synthesis of zeolite <span class="hlt">Na</span>A based on metakaolin were investigated according to results of cation exchange capacity and static water adsorption of all synthesis products and selected X-ray diffraction (XRD). Magnetic zeolite <span class="hlt">Na</span>A was synthesized by adding Fe3O4 in the precursor of zeolite. Zeolite <span class="hlt">Na</span>A and magnetic zeolite <span class="hlt">Na</span>A were characterized with scanning electron microscopy (SEM) and XRD. Magnetic zeolite <span class="hlt">Na</span>A with different Fe3O4 loadings was prepared and used for removal of heavy metals (Cu(<span class="hlt">2</span>+), Pb(<span class="hlt">2</span>+)). The results show the optimum parameters for synthesis zeolite <span class="hlt">Na</span>A are SiO<span class="hlt">2</span>/Al<span class="hlt">2</span>O3=<span class="hlt">2</span>.3, <span class="hlt">Na</span><span class="hlt">2</span>O/SiO<span class="hlt">2</span>=1.4, H<span class="hlt">2</span>O/<span class="hlt">Na</span><span class="hlt">2</span>O=50, crystallization time 8h, crystallization temperature 95 °C. The addition of Fe3O4 makes the <span class="hlt">Na</span>A zeolite with good magnetic susceptibility and good magnetic stability regardless of the Fe3O4 loading, confirming the considerable separation efficiency. Additionally, Fe3O4 loading had a little effect on removal of heavy metal by magnetic zeolite, however, the adsorption capacity still reaches <span class="hlt">2</span>.3 mmol g(-1) for Cu(<span class="hlt">2</span>+), Pb(<span class="hlt">2</span>+) with a removal efficiency of over 95% in spite of 4.7% Fe3O4 loading. This indicates magnetic zeolite can be used to remove metal heavy at least Cu(<span class="hlt">2</span>+), Pb(<span class="hlt">2</span>+) from water with metallic contaminants and can be separated easily after a magnetic process. Copyright © 2013 Elsevier Ltd. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013APS..MARC15011A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013APS..MARC15011A"><span>Frustration by competing interactions in the highly-distorted double perovskites La<span class="hlt">2</span><span class="hlt">Na</span>RuO6 and La<span class="hlt">2</span><span class="hlt">Na</span>OsO6</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Aczel, A. A.; Bugaris, D. E.; Li, L.; Yan, J.-Q.; de La Cruz, C.; Zur Loye, H.-C.; Nagler, S. E.</p> <p>2013-03-01</p> <p>The usual classical behavior of S = 3/<span class="hlt">2</span>, B-site ordered double perovskites results in simple, commensurate magnetic ground states. In contrast, heat capacity and neutron powder diffraction measurements for the S = 3/<span class="hlt">2</span> systems La<span class="hlt">2</span><span class="hlt">Na</span>B'O6 (B' = Ru, Os) reveal an incommensurate magnetic ground state for La<span class="hlt">2</span><span class="hlt">Na</span>RuO6 and a drastically suppressed ordered moment for La<span class="hlt">2</span><span class="hlt">Na</span>OsO6. This behavior is attributed to the large monoclinic structural distortions of these double perovskites. The distortions have the effect of weakening the nearest neighbor superexchange interactions, presumably to an energy scale that is comparable to the next nearest neighbor superexchange. The exotic ground states in these materials can then arise from a competition between these two types of antiferromagnetic interactions, providing a novel mechanism for achieving frustration in the double perovskite family. Work at ORNL is supported by the Division of Scientific User Facilities and the Materials Science and Engineering Division, DOE Basic Energy Sciences. Work at the University of South Carolina is supported by the Heterogeneous Functional Materials Research Center, funded by DOE under award number de-sc0001061.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4919009','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4919009"><span>Identification of a 3rd <span class="hlt">Na</span>+ Binding Site of the Glycine Transporter, GlyT<span class="hlt">2</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>Subramanian, Nandhitha; Scopelitti, Amanda J.; Carland, Jane E.; Ryan, Renae M.; O’Mara, Megan L.; Vandenberg, Robert J.</p> <p>2016-01-01</p> <p>The <span class="hlt">Na</span>+/Cl- dependent glycine transporters GlyT1 and GlyT<span class="hlt">2</span> regulate synaptic glycine concentrations. Glycine transport by GlyT<span class="hlt">2</span> is coupled to the co-transport of three <span class="hlt">Na</span>+ ions, whereas transport by GlyT1 is coupled to the co-transport of only two <span class="hlt">Na</span>+ ions. These differences in ion-flux coupling determine their respective concentrating capacities and have a direct bearing on their functional roles in synaptic transmission. The crystal structures of the closely related bacterial <span class="hlt">Na</span>+-dependent leucine transporter, LeuTAa, and the Drosophila dopamine transporter, dDAT, have allowed prediction of two <span class="hlt">Na</span>+ binding sites in GlyT<span class="hlt">2</span>, but the physical location of the third <span class="hlt">Na</span>+ site in GlyT<span class="hlt">2</span> is unknown. A bacterial betaine transporter, BetP, has also been crystallized and shows structural similarity to LeuTAa. Although betaine transport by BetP is coupled to the co-transport of two <span class="hlt">Na</span>+ ions, the first <span class="hlt">Na</span>+ site is not conserved between BetP and LeuTAa, the so called <span class="hlt">Na</span>1' site. We hypothesized that the third <span class="hlt">Na</span>+ binding site (<span class="hlt">Na</span>3 site) of GlyT<span class="hlt">2</span> corresponds to the BetP <span class="hlt">Na</span>1' binding site. To identify the <span class="hlt">Na</span>3 binding site of GlyT<span class="hlt">2</span>, we performed molecular dynamics (MD) simulations. Surprisingly, a <span class="hlt">Na</span>+ placed at the location consistent with the <span class="hlt">Na</span>1' site of BetP spontaneously dissociated from its initial location and bound instead to a novel <span class="hlt">Na</span>3 site. Using a combination of MD simulations of a comparative model of GlyT<span class="hlt">2</span> together with an analysis of the functional properties of wild type and mutant GlyTs we have identified an electrostatically favorable novel third <span class="hlt">Na</span>+ binding site in GlyT<span class="hlt">2</span> formed by Trp263 and Met276 in TM3, Ala481 in TM6 and Glu648 in TM10. PMID:27337045</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25537342','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25537342"><span>Distinctly Different Glass Transition Behaviors of Trehalose Mixed with <span class="hlt">Na</span><span class="hlt">2</span>HPO 4 or <span class="hlt">Na</span>H <span class="hlt">2</span>PO 4: Evidence for its Molecular Origin.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Weng, Lindong; Elliott, Gloria D</p> <p>2015-07-01</p> <p>The present study is aimed at understanding how the interactions between sugar molecules and phosphate ions affect the glass transition temperature of their mixtures, and the implications for pharmaceutical formulations. The glass transition temperature (Tg) and the α-relaxation temperature (Tα) of dehydrated trehalose/sodium phosphate mixtures (monobasic or dibasic) were determined by differential scanning calorimetry and dynamic mechanical analysis, respectively. Molecular dynamics simulations were also conducted to investigate the microscopic interactions between sugar molecules and phosphate ions. The hydrogen-bonding characteristics and the self-aggregation features of these mixtures were quantified and compared. Thermal analysis measurements demonstrated that the addition of <span class="hlt">Na</span>H<span class="hlt">2</span>PO4 decreased both the glass transition temperature and the α-relaxation temperature of the dehydrated trehalose/<span class="hlt">Na</span>H<span class="hlt">2</span>PO4 mixture compared to trehalose alone while both Tg and Tα were increased by adding <span class="hlt">Na</span><span class="hlt">2</span>HPO4 to pure trehalose. The hydrogen-bonding interactions between trehalose and HPO4(<span class="hlt">2</span>-) were found to be stronger than both the trehalose-trehalose hydrogen bonds and those formed between trehalose and H<span class="hlt">2</span>PO4(-). The HPO4(<span class="hlt">2</span>-) ions also aggregated into smaller clusters than H<span class="hlt">2</span>PO4(-) ions. The trehalose/<span class="hlt">Na</span><span class="hlt">2</span>HPO4 mixture yielded a higher T g than pure trehalose because marginally self-aggregated HPO4(<span class="hlt">2</span>-) ions established a strengthened hydrogen-bonding network with trehalose molecules. In contrast H<span class="hlt">2</span>PO4(-) ions served only as plasticizers, resulting in a lower Tg of the mixtures than trehalose alone, creating large-sized ionic pockets, weakening interactions, and disrupting the original hydrogen-bonding network amongst trehalose molecules.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28822959','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28822959"><span><span class="hlt">Na</span><span class="hlt">2</span>S, a fast-releasing H<span class="hlt">2</span>S donor, given as suppository lowers blood pressure in rats.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Tomasova, Lenka; Drapala, Adrian; Jurkowska, Halina; Wróbel, Maria; Ufnal, Marcin</p> <p>2017-10-01</p> <p>Hydrogen sulfide (H <span class="hlt">2</span> S) is involved in blood pressure control. The available slow-releasing H <span class="hlt">2</span> S-donors are poorly soluble in water and their ability to release H <span class="hlt">2</span> S in biologically relevant amounts under physiological conditions is questionable. Therefore, new slow-releasing donors or new experimental approaches to fast-releasing H <span class="hlt">2</span> S donors are needed. Hemodynamics and ECG were recorded in male, anesthetized Wistar Kyoto rats (WKY) and in Spontaneously hypertensive rats (SHR) at baseline and after: 1) intravenous (iv) infusion of vehicle or <span class="hlt">Na</span> <span class="hlt">2</span> S; <span class="hlt">2</span>) administration of vehicle suppositories or <span class="hlt">Na</span> <span class="hlt">2</span> S suppositories. Intravenously administered vehicle and vehicle suppositories did not affect mean arterial blood pressure (MABP) and heart rate (HR). <span class="hlt">Na</span> <span class="hlt">2</span> S administered iv caused a significant, but transient (<span class="hlt">2</span>-5min) decrease in MABP. <span class="hlt">Na</span> <span class="hlt">2</span> S suppositories produced a dose-dependent hypotensive response that lasted ∼45min in WKY and ∼75-80min in SHR. It was accompanied by a decrease in HR in WKY, and an increase in HR in SHR. <span class="hlt">Na</span> <span class="hlt">2</span> S suppositories did not produce a significant change in corrected QT, an indicator of cardiotoxicity. <span class="hlt">Na</span> <span class="hlt">2</span> S suppositories increased blood level of thiosulfates, products of H <span class="hlt">2</span> S oxidation. <span class="hlt">Na</span> <span class="hlt">2</span> S administered in suppositories exerts a prolonged hypotensive effect in rats, with no apparent cardiotoxic effect. SHR and WKY differ in hemodynamic response to the H <span class="hlt">2</span> S donor. Suppository formulation of fast-releasing H <span class="hlt">2</span> S donors may be useful in research, if a reference slow-releasing H <span class="hlt">2</span> S donor is not available. Copyright © 2017 Institute of Pharmacology, Polish Academy of Sciences. Published by Elsevier Urban & Partner Sp. z o.o. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1226909-surface-naau2-structure-composition-stability','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1226909-surface-naau2-structure-composition-stability"><span>The (111) Surface of <span class="hlt">Na</span>Au <span class="hlt">2</span>. Structure, Composition, and Stability</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Kwolek, Emma J.; Widmer, Roland; Gröning, Oliver; ...</p> <p>2014-12-17</p> <p>The (111) surface of single-crystal <span class="hlt">Na</span>Au <span class="hlt">2</span> is a model for catalytically active, powdered <span class="hlt">Na</span>Au <span class="hlt">2</span>. We prepare and characterize this surface with a broad suite of techniques. Preparation in ultrahigh vacuum consists of the traditional approach of ion bombardment (to remove impurities) and thermal annealing (to restore surface order). Both of these steps cause loss of sodium (<span class="hlt">Na</span>), and repeated treatments eventually trigger conversion of the surface and near-surface regions to crystalline gold. The bulk has a limited ability to repopulate the surface <span class="hlt">Na</span>. Under conditions where <span class="hlt">Na</span> depletion is minimized, electron diffraction patterns are consistent with the bulk-terminatedmore » structure, and scanning tunneling microscopy reveals mesa-like features with lateral dimensions of a few tens of nanometers. The tops of the mesas do not possess fine structure characteristic of a periodic lattice, suggesting that the surface layer is disordered under the conditions of these experiments.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23869994','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23869994"><span>Growth responses and ion accumulation in the halophytic legume Prosopis strombulifera are determined by <span class="hlt">Na</span><span class="hlt">2</span>SO4 and <span class="hlt">Na</span>Cl.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Reginato, M; Sosa, L; Llanes, A; Hampp, E; Vettorazzi, N; Reinoso, H; Luna, V</p> <p>2014-01-01</p> <p>Halophytes are potential gene sources for genetic manipulation of economically important crop species. This study addresses the physiological responses of a widespread halophyte, Prosopis strombulifera (Lam.) Benth to salinity. We hypothesised that increasing concentrations of the two major salts present in soils of central Argentina (<span class="hlt">Na</span><span class="hlt">2</span>SO4, <span class="hlt">Na</span>Cl, or their iso-osmotic mixture) would produce distinct physiological responses. We used hydroponically grown P. strombulifera to test this hypothesis, analysing growth parameters, water relations, photosynthetic pigments, cations and anions. These plants showed a halophytic response to <span class="hlt">Na</span>Cl, but strong general inhibition of growth in response to iso-osmotic solutions containing <span class="hlt">Na</span><span class="hlt">2</span>SO4. The explanation for the adaptive success of P. strombulifera in high <span class="hlt">Na</span>Cl conditions seems to be related to a delicate balance between <span class="hlt">Na</span>(+) accumulation (and its use for osmotic adjustment) and efficient compartmentalisation in vacuoles, the ability of the whole plant to ensure sufficient K(+) supply by maintaining high K(+)/<span class="hlt">Na</span>(+) discrimination, and maintenance of normal Ca(<span class="hlt">2</span>+) levels in leaves. The three salt treatments had different effects on the accumulation of ions. Findings in bi-saline-treated plants were of particular interest, where most of the physiological parameters studied showed partial alleviation of SO4(<span class="hlt">2</span>-)-induced toxicity by Cl(-). Thus, discussions on physiological responses to salinity could be further expanded in a way that more closely mimics natural salt environments. © 2013 German Botanical Society and The Royal Botanical Society of the Netherlands.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5455555','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5455555"><span>Effect of Solids-To-Liquids, <span class="hlt">Na</span><span class="hlt">2</span>SiO3-To-<span class="hlt">Na</span>OH and Curing Temperature on the Palm Oil Boiler Ash (Si + Ca) Geopolymerisation System</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Yahya, Zarina; Abdullah, Mohd Mustafa Al Bakri; Hussin, Kamarudin; Ismail, Khairul Nizar; Abd Razak, Rafiza; Sandu, Andrei Victor</p> <p>2015-01-01</p> <p>This paper investigates the effect of the solids-to-liquids (S/L) and <span class="hlt">Na</span><span class="hlt">2</span>SiO3/<span class="hlt">Na</span>OH ratios on the production of palm oil boiler ash (POBA) based geopolymer. Sodium silicate and sodium hydroxide (<span class="hlt">Na</span>OH) solution were used as alkaline activator with a <span class="hlt">Na</span>OH concentration of 14 M. The geopolymer samples were prepared with different S/L ratios (0.5, 1.0, 1.25, 1.5, and 1.75) and <span class="hlt">Na</span><span class="hlt">2</span>SiO3/<span class="hlt">Na</span>OH ratios (0.5, 1.0, 1.5, <span class="hlt">2</span>.0, <span class="hlt">2</span>.5, and 3.0). The main evaluation techniques in this study were compressive strength, X-Ray Diffraction (XRD), Fourier Transform Infrared Spectroscopy (FTIR), and Scanning Electron Microscope (SEM). The results showed that the maximum compressive strength (11.9 MPa) was obtained at a S/L ratio and <span class="hlt">Na</span><span class="hlt">2</span>SiO3/<span class="hlt">Na</span>OH ratio of 1.5 and <span class="hlt">2</span>.5 at seven days of testing.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2004JSSCh.177.4475E','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2004JSSCh.177.4475E"><span>Synthesis and characterization of a <span class="hlt">Na</span>SICON series with general formula <span class="hlt">Na</span> <span class="hlt">2</span>.8Zr <span class="hlt">2</span>-ySi 1.8-4yP 1.<span class="hlt">2</span>+4yO 12 (0⩽ y⩽0.45)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Essoumhi, A.; Favotto, C.; Mansori, M.; Satre, P.</p> <p>2004-12-01</p> <p>In this work, we present the synthesis and the characterization of ionic conducting ceramics of <span class="hlt">Na</span>SICON-type (Natrium super ionic conductor). The properties of this ceramic make it suitable for use in electrochemical devices. These solid electrolytes can be used as sensors for application in the manufacturing of potentiometric gas sensors, for the detection of pollutant emissions and for environment control. The family of <span class="hlt">Na</span>SICON that we studied has as a general formula <span class="hlt">Na</span> <span class="hlt">2</span>.8Zr <span class="hlt">2</span>-ySi 1.8-4yP 1.<span class="hlt">2</span>+4yO 12 with 0⩽ y⩽0.45. The various compositions were synthesized by produced using the sol-gel method. The electric properties of these compositions were carried out by impedance spectroscopy. The results highlight the good conductivity of the <span class="hlt">Na</span> <span class="hlt">2</span>.8Zr 1.775Si 0.9P <span class="hlt">2</span>.1O 12 composition.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26461467','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26461467"><span><span class="hlt">2</span>D Electrides as Promising Anode Materials for <span class="hlt">Na</span>-Ion Batteries from First-Principles Study.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Hu, Junping; Xu, Bo; Yang, Shengyuan A; Guan, Shan; Ouyang, Chuying; Yao, Yugui</p> <p>2015-11-04</p> <p>Searching for suitable anodes with good performance is a key challenge for rechargeable <span class="hlt">Na</span>-ion batteries (NIBs). Using the first-principles method, we predict that <span class="hlt">2</span>D nitrogen electride materials can be served as anode materials for NIBs. Particularly, we show that Ca<span class="hlt">2</span>N meets almost all the requirements of a good NIB anode. Each formula unit of a monolayer Ca<span class="hlt">2</span>N sheet can absorb up to four <span class="hlt">Na</span> atoms, corresponding to a theoretical specific capacity of 1138 mAh·g(-1). The metallic character for both pristine Ca<span class="hlt">2</span>N and its <span class="hlt">Na</span> intercalated state NaxCa<span class="hlt">2</span>N ensures good electronic conduction. <span class="hlt">Na</span> diffusion along the <span class="hlt">2</span>D monolayer plane can be very fast even at room temperature, with a <span class="hlt">Na</span> migration energy barrier as small as 0.084 eV. These properties are key to the excellent rate performance of an anode material. The average open-circuit voltage is calculated to be 0.18 V vs <span class="hlt">Na/Na</span>(+) for the chemical stoichiometry of <span class="hlt">Na</span><span class="hlt">2</span>Ca<span class="hlt">2</span>N and 0.09 V for <span class="hlt">Na</span>4Ca<span class="hlt">2</span>N. The relatively low average open-circuit voltage is beneficial to the overall voltage of the cell. In addition, the <span class="hlt">2</span>D monolayers have very small lattice change upon <span class="hlt">Na</span> intercalation, which ensures a good cycling stability. All these results demonstrate that the Ca<span class="hlt">2</span>N monolayer could be an excellent anode material for NIBs.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018RJPCA..92..475C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018RJPCA..92..475C"><span>Phase Diagram of Quaternary System <span class="hlt">Na</span>Br-KBr-CaBr<span class="hlt">2</span>-H<span class="hlt">2</span>O at 323 K</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Cui, Rui-Zhi; Wang, Wei; Yang, Lei; Sang, Shi-Hua</p> <p>2018-03-01</p> <p>The phase equilibria in the system <span class="hlt">Na</span>Br-KBr-CaBr<span class="hlt">2</span>-H<span class="hlt">2</span>O at 323 K were studied using the isothermal dissolution equilibrium method. Using the experimental solubilities of salts data, phase diagram was constructed. The phase diagram have two invariant points, five univariant curves, and four crystallization fields. The equilibrium solid phases in the system are <span class="hlt">Na</span>Br, <span class="hlt">Na</span>Br · <span class="hlt">2</span>H<span class="hlt">2</span>O, KBr, and CaBr<span class="hlt">2</span> · 4H<span class="hlt">2</span>O. The solubilities of salts in the system at 323 K were calculated by Pitzer's equation. There is shown that the calculated solubilities agree well with experimental data.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JPS...370..114L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JPS...370..114L"><span>Structural and electrochemical properties of Fe-doped <span class="hlt">Na</span><span class="hlt">2</span>Mn3-xFex(P<span class="hlt">2</span>O7)<span class="hlt">2</span> cathode material for sodium ion batteries</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Liu, Huatao; Zhao, Yanming; Zhang, Hui; Lian, Xin; Dong, Youzhong; Kuang, Quan</p> <p>2017-12-01</p> <p>A series of Fe-doped <span class="hlt">Na</span><span class="hlt">2</span>Mn3-xFex(P<span class="hlt">2</span>O7)<span class="hlt">2</span> (x = 0.0, 0.5, 1.0, 1.5 and <span class="hlt">2</span>.0) compounds have been successfully prepared by using sol-gel method. Rietveld refinement results indicate that single phase <span class="hlt">Na</span><span class="hlt">2</span>Mn3-xFex(P<span class="hlt">2</span>O7)<span class="hlt">2</span> with triclinic structure can be obtained within 0 ≤ x ≤ <span class="hlt">2</span> although no <span class="hlt">Na</span><span class="hlt">2</span>Fe3(P<span class="hlt">2</span>O7)<span class="hlt">2</span> existing under our experimental conditions, and the cell parameters (including a, b, c and V) are decreasing with the increasing of x. Our results reveal that <span class="hlt">Na</span><span class="hlt">2</span>Mn3(P<span class="hlt">2</span>O7)<span class="hlt">2</span> exhibits an electrochemical activity in the voltage range of 1.5 V-4.5 V vs. <span class="hlt">Na+/Na</span> when using as the cathode material for SIBs although it gives a limited rate capability and poor capacity retention. However, the electrochemical performance of Fe-doped <span class="hlt">Na</span><span class="hlt">2</span>Mn3-xFex(P<span class="hlt">2</span>O7)<span class="hlt">2</span> (0 ≤ x ≤ <span class="hlt">2</span>) can be improved significantly where cycle performance and rate capability can be improved significantly than that of the pristine one. Sodium ion diffusion coefficient can be increased by about two orders of magnitude with the Fe-doping content higher than x = 0.5.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2003JSSCh.172...95R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2003JSSCh.172...95R"><span>The crystal structure of synthetic simmonsite, <span class="hlt">Na</span> <span class="hlt">2</span>LiAlF 6</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ross, Kirk C.; Mitchell, Roger H.; Chakhmouradian, Anton R.</p> <p>2003-04-01</p> <p>The structure of the synthetic fluoroperovskite, <span class="hlt">Na</span> <span class="hlt">2</span>LiAlF 6 (simmonsite), has been determined by powder X-ray diffraction using the Rietveld method of structure refinement. The compound adopts space group P<span class="hlt">2</span> 1/ n [#14; a=5.2842(1); b=5.3698(1); c=7.5063(<span class="hlt">2</span>) Å; β=89.98(1)°; Z=4), and is a member of the cryolite (<span class="hlt">Na</span> <span class="hlt">2</span><span class="hlt">Na</span>AlF 6) structural group characterized by ordering of the B-site cations (Li, Al) and tilting of the BF 6 octahedra according to the tilt scheme a-b-c+. Rotations of the B-site polyhedra are less ( ΦLi=14.9°; ΦAl=17.0°) than those found in cryolite ( Φ<span class="hlt">Na</span>=18.6; ΦAl=23.5°) because of the larger difference in the ionic radii of the B-site cations in cryolite as compared to those in simmonsite. <span class="hlt">Na</span> at the A-site is displaced from the special position resulting in 10- and 8-fold coordination in simmonsite and cryolite, respectively. By analogy with the synthetic compound, naturally occurring simmonsite is considered to adopt space group P<span class="hlt">2</span> 1/ n (#14) and not the P<span class="hlt">2</span> 1(#4) or P<span class="hlt">2</span> 1/ m(#11) space groups.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015PhSS...57.1198C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015PhSS...57.1198C"><span>Lattice dynamics of Cs<span class="hlt">2</span><span class="hlt">Na</span>YbF6 and Cs<span class="hlt">2</span><span class="hlt">Na</span>YF6 elpasolites: Ab initio calculation</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Chernyshev, V. A.; Petrov, V. P.; Nikiforov, A. E.; Zakir'yanov, D. O.</p> <p>2015-06-01</p> <p>The ab initio calculations of the crystal structure and the phonon spectrum of Cs<span class="hlt">2</span><span class="hlt">Na</span>YbF6 and Cs<span class="hlt">2</span><span class="hlt">Na</span>YF6 crystals with the elpasolite structure have been performed. The frequencies and types of fundamental vibrations have been determined. The calculations have been performed in the framework of the density functional theory using the molecular orbital method with hybrid functionals in the CRYSTAL09 program developed for the simulation of periodic structures. The outer 5 s and 5 p shells of the rare-earth ion have been described in Gaussian-type basis sets. The influence of inner shells, including 4 f electron shells, on the outer shells has been described using the pseudopotential. It has been shown that this approach allows the description of the phonon spectrum with the inclusion of the splitting of the longitudinal and transverse optical modes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018EML...tmp..112S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018EML...tmp..112S"><span>Optical and Luminescence Properties of β-<span class="hlt">Na</span>FeO<span class="hlt">2</span> Nanoparticles</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Singh, Sarbjit; Tangra, Ankush Kumar; Lotey, Gurmeet Singh</p> <p>2018-05-01</p> <p>β-<span class="hlt">Na</span>FeO<span class="hlt">2</span> nanoparticles have been synthesized by sol-gel method and their morphological, structural and optical properties investigated. Transmission electron microscope study reveals that the size of the synthesis nanoparticles is 37 nm and they are possessing spherical symmetry. X-ray diffraction pattern shows the orthorhombic crystal structure of nanoparticles with space group Pn21 a. UV-visible spectra of β-<span class="hlt">Na</span>FeO<span class="hlt">2</span> divulges that these nanoparticles have direct band gap <span class="hlt">2</span>.35 eV. The observed Fourier transform infrared spectroscopy spectra confirms the presence of Fe-<span class="hlt">Na</span> bonding at 1074 cm-1. The photoluminescence study of these nanoparticles shows that these nanoparticles possesses various transition in the visible spectrum.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29226609','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29226609"><span>A P<span class="hlt">2</span>-Type Layered Superionic Conductor Ga-Doped <span class="hlt">Na</span><span class="hlt">2</span> Zn<span class="hlt">2</span> TeO6 for All-Solid-State Sodium-Ion Batteries.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Li, Yuyu; Deng, Zhi; Peng, Jian; Chen, Enyi; Yu, Yao; Li, Xiang; Luo, Jiahuan; Huang, Yangyang; Zhu, Jinlong; Fang, Chun; Li, Qing; Han, Jiantao; Huang, Yunhui</p> <p>2018-01-24</p> <p>Here, a P<span class="hlt">2</span>-type layered <span class="hlt">Na</span> <span class="hlt">2</span> Zn <span class="hlt">2</span> TeO 6 (NZTO) is reported with a high <span class="hlt">Na</span> + ion conductivity ≈0.6×10 -3  S cm -1 at room temperature (RT), which is comparable to the currently best <span class="hlt">Na</span> 1+n Zr <span class="hlt">2</span> Si n P 3-n O 12 NASICON structure. As small amounts of Ga 3+ substitutes for Zn <span class="hlt">2</span>+ , more <span class="hlt">Na</span> + vacancies are introduced in the interlayer gaps, which greatly reduces strong <span class="hlt">Na</span> + -<span class="hlt">Na</span> + coulomb interactions. Ga-substituted NZTO exhibits a superionic conductivity of ≈1.1×10 -3  S cm -1 at RT, and excellent phase and electrochemical stability. All solid-state batteries have been successfully assembled with a capacity of ≈70 mAh g -1 over 10 cycles with a rate of 0.<span class="hlt">2</span> C at 80 °C. 23 <span class="hlt">Na</span> nuclear magnetic resonance (NMR) studies on powder samples show intra-grain (bulk) diffusion coefficients D NMR on the order of 12.35×10 -12  m <span class="hlt">2</span>  s -1 at 65 °C that corresponds to a conductivity σ NMR of 8.16×10 -3  S cm -1 , assuming the Nernst-Einstein equation, which thus suggests a new perspective of fast <span class="hlt">Na</span> + ion conductor for advanced sodium ion batteries. © 2018 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29283577','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29283577"><span>Reduction Mechanisms of Cu<span class="hlt">2</span>+-Doped <span class="hlt">Na</span><span class="hlt">2</span>O-Al<span class="hlt">2</span>O3-SiO<span class="hlt">2</span> Glasses during Heating in H<span class="hlt">2</span> Gas.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Nogami, Masayuki; Quang, Vu Xuan; Ohki, Shinobu; Deguchi, Kenzo; Shimizu, Tadashi</p> <p>2018-01-25</p> <p>Controlling valence state of metal ions that are doped in materials has been widely applied for turning optical properties. Even though hydrogen has been proven effective to reduce metal ions because of its strong reducing capability, few comprehensive studies focus on practical applications because of the low diffusion rate of hydrogen in solids and the limited reaction near sample surfaces. Here, we investigated the reactions of hydrogen with Cu <span class="hlt">2</span>+ -doped <span class="hlt">Na</span> <span class="hlt">2</span> O-Al <span class="hlt">2</span> O 3 -SiO <span class="hlt">2</span> glass and found that a completely different reduction from results reported so far occurs, which is dominated by the Al/<span class="hlt">Na</span> concentration ratio. For Al/<span class="hlt">Na</span> < 1, Cu <span class="hlt">2</span>+ ions were reduced via hydrogen to metallic Cu, distributing in glass body. For Al/<span class="hlt">Na</span> > 1, on the other hand, the reduction of Cu <span class="hlt">2</span>+ ions occurred simultaneously with the formation of OH bonds, whereas the reduced Cu metal moved outward and formed a metallic film on glass surface. The NMR and Fourier transform infrared results indicated that the Cu <span class="hlt">2</span>+ ions were surrounded by Al 3+ ions that formed AlO 4 , distorted AlO 4 , and AlO 5 units. The diffused H <span class="hlt">2</span> gas reacted with the Al-O - ···Cu + units, forming Al-OH and metallic Cu, the latter of which moved freely toward glass surface and in return enhanced H <span class="hlt">2</span> diffusion.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018ApSS..433..341X','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018ApSS..433..341X"><span>Effect of <span class="hlt">Na</span> poisoning catalyst (V<span class="hlt">2</span>O5-WO3/TiO<span class="hlt">2</span>) on denitration process and SO3 formation</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Xiao, Haiping; Chen, Yu; Qi, Cong; Ru, Yu</p> <p>2018-03-01</p> <p>This paper aims to study the effect of alkali metal sodium (<span class="hlt">Na</span>) poisoning on the performance of the Selective Catalytic Reduction (SCR) catalyst. The result showed that <span class="hlt">Na</span><span class="hlt">2</span>SO4 poisoning leads to a reduced denitration rate of the SCR catalyst and an increase in the SO3 generation rate. <span class="hlt">Na</span><span class="hlt">2</span>O poisoning leads to a significant reduction in the denitration rate of the SCR catalyst and marginally improves the formation of SO3. The maximum of the SO3 generation rate for a <span class="hlt">Na</span><span class="hlt">2</span>SO4-poisoned catalyst reached 1.35%, whereas it was only 0.85% for the SCR catalyst. When the SO<span class="hlt">2</span> was contained in flue gas, the denitration rate for the <span class="hlt">Na</span><span class="hlt">2</span>O-poisoned catalyst clearly increased by more than 28%. However, the effect of SO<span class="hlt">2</span> on the <span class="hlt">Na</span><span class="hlt">2</span>SO4-poisoned catalyst was very slight. The denitration rate of the SCR catalyst decreased with an increase in the <span class="hlt">Na</span> content. The BET and XRD results showed that <span class="hlt">Na</span> poisoning of the catalyst decreased the number of acid sites, the reducibility of the catalyst, the surface area, and pore volume. The H<span class="hlt">2</span>-TPR and NH3-TPD results show that <span class="hlt">Na</span> decreases the number of acid sites and the reducibility of the catalyst. The FT-IR and XPS results showed that <span class="hlt">Na</span><span class="hlt">2</span>O poisoning led to the decrease of V5+dbnd O bonds and the consumptions of oxygen atoms. <span class="hlt">Na</span><span class="hlt">2</span>SO4 poisoning can improve surface adsorbed oxygen, which was beneficial for the SO<span class="hlt">2</span>-SO3 conversion reaction.</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/2015RJPCA..89..634R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015RJPCA..89..634R"><span>Phase diagram of the LiNO3-<span class="hlt">Na</span>NO3-<span class="hlt">Na</span>Cl-Sr(NO3)<span class="hlt">2</span> salt system</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Rasulov, A. I.; Gasanaliev, A. M.; Mamedova, A. K.; Gamataeva, B. Yu.</p> <p>2015-04-01</p> <p>The phase diagram of the quaternary LiNO3-<span class="hlt">Na</span>NO3-<span class="hlt">Na</span>Cl-Sr(NO3)<span class="hlt">2</span> system is studied by means of differential thermal analysis, and the compositions and crystallization temperatures of nonvariant equilibrium phases are revealed. The temperature dependence of conductivity in eutectic and peritectic salt compositions is investigated.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29493210','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29493210"><span>3.0 V High Energy Density Symmetric Sodium-Ion Battery: <span class="hlt">Na</span>4V<span class="hlt">2</span>(PO4)3∥<span class="hlt">Na</span>3V<span class="hlt">2</span>(PO4)3.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Yao, Xuhui; Zhu, Zixuan; Li, Qi; Wang, Xuanpeng; Xu, Xiaoming; Meng, Jiashen; Ren, Wenhao; Zhang, Xinhe; Huang, Yunhui; Mai, Liqiang</p> <p>2018-03-28</p> <p>Symmetric sodium-ion batteries (SIBs) are considered as promising candidates for large-scale energy storage owing to the simplified manufacture and wide abundance of sodium resources. However, most symmetric SIBs suffer from suppressed energy density. Here, a superior congeneric <span class="hlt">Na</span> 4 V <span class="hlt">2</span> (PO 4 ) 3 anode is synthesized via electrochemical preintercalation, and a high energy density symmetric SIB (<span class="hlt">Na</span> 3 V <span class="hlt">2</span> (PO 4 ) 3 as a cathode and <span class="hlt">Na</span> 4 V <span class="hlt">2</span> (PO 4 ) 3 as an anode) based on the deepened redox couple of V 4+ /V <span class="hlt">2</span>+ is built for the first time. When measured in half cell, both electrodes show stabilized electrochemical performance (over 3000 cycles). The symmetric SIBs exhibit an output voltage of 3.0 V and a cell-level energy density of 138 W h kg -1 . Furthermore, the sodium storage mechanism under the expanded measurement range of 0.01-3.9 V is disclosed through an in situ X-ray diffraction technique.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/1415478-na-ion-intercalation-charge-storage-mechanism-vanadium-carbide','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/1415478-na-ion-intercalation-charge-storage-mechanism-vanadium-carbide"><span><span class="hlt">Na</span>-Ion Intercalation and Charge Storage Mechanism in <span class="hlt">2</span>D Vanadium Carbide</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>Bak, Seong-Min; Qiao, Ruimin; Yang, Wanli</p> <p></p> <p>Two-dimensional vanadium carbide MXene containing surface functional groups (denoted as V<span class="hlt">2</span>CTx, where Tx are surface functional groups) was synthesized and studied as anode material for <span class="hlt">Na</span>-ion batteries. V<span class="hlt">2</span>CTx anode exhibits reversible charge storage with good cycling stability and high rate capability through electrochemical test. The charge storage mechanism of V<span class="hlt">2</span>CTx material during <span class="hlt">Na</span>+ intercalation/deintercalation and the redox reaction of vanadium were studied using a combination of synchrotron based X-ray diffraction (XRD), hard X-ray absorption near edge spectroscopy (XANES) and soft X-ray absorption spectroscopy (sXAS). Experimental evidence of a major contribution of redox reaction of vanadium to the charge storage andmore » the reversible capacity of V<span class="hlt">2</span>CTx during sodiation/desodiation process have been provided through V K-edge XANES and V L<span class="hlt">2</span>,3-edge sXAS results. A correlation between the CO32- content and <span class="hlt">Na</span>+ intercalation/deintercalation states in the V<span class="hlt">2</span>CTx electrode observed from C and O K-edge in sXAS results imply that some additional charge storage reactions may take place between the <span class="hlt">Na</span>+-intercalated V<span class="hlt">2</span>CTx and the carbonate based non-aqueous electrolyte. The results of this study will provide valuable information for the further studies on V<span class="hlt">2</span>CTx as anode material for <span class="hlt">Na</span>-ion batteries and capacitors.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29873483','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29873483"><span>Dithallium(III)-Containing 30-Tungsto-4-phosphate, [Tl<span class="hlt">2</span><span class="hlt">Na</span><span class="hlt">2</span>(H<span class="hlt">2</span>O)<span class="hlt">2</span>(P<span class="hlt">2</span>W15O56)<span class="hlt">2</span>]16-: Synthesis, Structural Characterization, and Biological Studies.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Ayass, Wassim W; Fodor, Tamás; Farkas, Edit; Lin, Zhengguo; Qasim, Hafiz M; Bhattacharya, Saurav; Mougharbel, Ali S; Abdallah, Khaled; Ullrich, Matthias S; Zaib, Sumera; Iqbal, Jamshed; Harangi, Sándor; Szalontai, Gábor; Bányai, István; Zékány, László; Tóth, Imre; Kortz, Ulrich</p> <p>2018-06-18</p> <p>Here we report on the synthesis and structural characterization of the dithallium(III)-containing 30-tungsto -4-phosphate [Tl <span class="hlt">2</span> <span class="hlt">Na</span> <span class="hlt">2</span> (H <span class="hlt">2</span> O) <span class="hlt">2</span> {P <span class="hlt">2</span> W 15 O 56 } <span class="hlt">2</span> ] 16- (1) by a multitude of solid-state and solution techniques. Polyanion 1 comprises two octahedrally coordinated Tl 3+ ions sandwiched between two trilacunary {P <span class="hlt">2</span> W 15 } Wells-Dawson fragments and represents only the second structurally characterized, discrete thallium-containing polyoxometalate to date. The two outer positions of the central rhombus are occupied by sodium ions. The title polyanion is solution-stable as shown by 31 P and 203/205 Tl NMR. This was also supported by Tl NMR spectra simulations including several spin systems of isotopologues with half-spin nuclei ( 203 Tl, 205 Tl, 31 P, 183 W). 23 <span class="hlt">Na</span> NMR showed a time-averaged signal of the <span class="hlt">Na</span> + counter cations and the structurally bonded <span class="hlt">Na</span> + ions. 203/205 Tl NMR spectra also showed a minor signal tentatively attributed to the trithallium-containing derivative [Tl 3 <span class="hlt">Na</span>(H <span class="hlt">2</span> O) <span class="hlt">2</span> (P <span class="hlt">2</span> W 15 O 56 ) <span class="hlt">2</span> ] 14- , which could also be identified in the solid state by single-crystal X-ray diffraction. The bioactivity of polyanion 1 was also tested against bacteria and Leishmania.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4509643','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4509643"><span>Distinctly Different Glass Transition Behaviors of Trehalose Mixed with <span class="hlt">Na</span><span class="hlt">2</span>HPO4 or <span class="hlt">Na</span>H<span class="hlt">2</span>PO4: Evidence for its Molecular Origin</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Weng, Lindong; Elliott, Gloria D.</p> <p>2015-01-01</p> <p>Purpose The present study is aimed at understanding how the interactions between sugar molecules and phosphate ions affect the glass transition temperature of their mixtures, and the implications for pharmaceutical formulations. Methods The glass transition temperature (Tg) and the α-relaxation temperature (Tα) of dehydrated trehalose/sodium phosphate mixtures (monobasic or dibasic) were determined by differential scanning calorimetry and dynamic mechanical analysis, respectively. Molecular dynamics simulations were also conducted to investigate the microscopic interactions between sugar molecules and phosphate ions. The hydrogen-bonding characteristics and the self-aggregation features of these mixtures were quantified and compared. Results Thermal analysis measurements demonstrated that the addition of <span class="hlt">Na</span>H<span class="hlt">2</span>PO4 decreased both the glass transition temperature and the α-relaxation temperature of the dehydrated trehalose/<span class="hlt">Na</span>H<span class="hlt">2</span>PO4 mixture compared to trehalose alone while both Tg and Tα were increased by adding <span class="hlt">Na</span><span class="hlt">2</span>HPO4 to pure trehalose. The hydrogen-bonding interactions between trehalose and HPO42− were found to be stronger than both the trehalose-trehalose hydrogen bonds and those formed between trehalose and H<span class="hlt">2</span>PO4−. The HPO42− ions also aggregated into smaller clusters than H<span class="hlt">2</span>PO4− ions. Conclusions The trehalose/<span class="hlt">Na</span><span class="hlt">2</span>HPO4 mixture yielded a higher Tg than pure trehalose because marginally self-aggregated HPO42− ions established a strengthened hydrogen-bonding network with trehalose molecules. In contrast H<span class="hlt">2</span>PO4− ions served only as plasticizers, resulting in a lower Tg of the mixtures than trehalose alone, creating large-sized ionic pockets, weakening interactions, and disrupting the original hydrogen-bonding network amongst trehalose molecules. PMID:25537342</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1152620','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1152620"><span>Adrenergic-agonist-induced Ca<span class="hlt">2</span>+ fluxes in rat parotid cells are not <span class="hlt">Na</span>+-dependent.</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Helman, J; Roth, G S; Baum, B J</p> <p>1985-01-01</p> <p>We investigated the hypothesis that extracellular <span class="hlt">Na</span>+ is required for the rapid mobilization of Ca<span class="hlt">2</span>+ by rat parotid cells after adrenergic stimulation. When <span class="hlt">Na</span>+ salts in the media were osmotically replaced with either choline chloride (+atropine) or sucrose, efflux of 45Ca<span class="hlt">2</span>+ from preloaded cells, caused by 10 microM-(-)-adrenaline, was unchanged. Similarly adrenaline stimulated 45Ca<span class="hlt">2</span>+ uptake into cells under nonsteady-state conditions in the presence or absence of <span class="hlt">Na</span>+. Monensin, a <span class="hlt">Na</span>+ ionophore, was able to elicit a modest increase in 45Ca<span class="hlt">2</span>+ efflux, compared with controls. Studies of net 45Ca<span class="hlt">2</span>+ flux, performed under near-steady-state conditions, showed that adrenaline caused net 45Ca<span class="hlt">2</span>+ accumulation, whereas monensin caused net 45Ca<span class="hlt">2</span>+ release. The effect of monensin required the presence of <span class="hlt">Na</span>+ in the incubation medium. Both 1 mM-LaCl3 and 0.1 mM-D-600 prevented adrenaline-stimulated 45Ca<span class="hlt">2</span>+ uptake into cells, but had no effect on monensin-induced changes. We conclude that (1) the rapid mobilization of Ca<span class="hlt">2</span>+ by adrenergic agonists seen in rat parotid cells does not require a <span class="hlt">Na</span>+out greater than <span class="hlt">Na</span>+in gradient and (<span class="hlt">2</span>) the nature of the monensin effect is quite different from the adrenergic-agonist-induced response. PMID:2413840</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/21842933','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21842933"><span>The study for the incipient solvation process of <span class="hlt">Na</span>Cl in water: the observation of the <span class="hlt">Na</span>Cl-(H<span class="hlt">2</span>O)n (n = 1, <span class="hlt">2</span>, and 3) complexes using Fourier-transform microwave spectroscopy.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Mizoguchi, Asao; Ohshima, Yasuhiro; Endo, Yasuki</p> <p>2011-08-14</p> <p>Pure rotational spectra of the sodium chloride-water complexes, <span class="hlt">Na</span>Cl-(H(<span class="hlt">2</span>)O)(n) (n = 1, <span class="hlt">2</span>, and 3), in the vibronic ground state have been observed by a Fourier- transform microwave spectrometer coupled with a laser ablation source. The (37)Cl-isotopic species and a few deuterated species have also been observed. From the analyses of the spectra, the rotational constants, the centrifugal distortion constants, and the nuclear quadrupole coupling constants of the <span class="hlt">Na</span> and Cl nuclei were determined precisely for all the species. The molecular structures of <span class="hlt">Na</span>Cl-(H(<span class="hlt">2</span>)O)(n) were determined using the rotational constants and the molecular symmetry. The charge distributions around <span class="hlt">Na</span> and Cl nuclei in <span class="hlt">Na</span>Cl are dramatically changed by the complex formation with H(<span class="hlt">2</span>)O. Prominent dependences of the bond lengths r(<span class="hlt">Na</span>-Cl) on the number of H(<span class="hlt">2</span>)O were also observed. By a comparison with results of theoretical studies, it is shown that the structure of <span class="hlt">Na</span>Cl-(H(<span class="hlt">2</span>)O)(3) is approaching to that of the contact ion-pair, which is considered to be an intermediate species in the incipient solvation process.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25468964','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25468964"><span>Sodium recognition by the <span class="hlt">Na</span>+/Ca<span class="hlt">2</span>+ exchanger in the outward-facing conformation.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Marinelli, Fabrizio; Almagor, Lior; Hiller, Reuben; Giladi, Moshe; Khananshvili, Daniel; Faraldo-Gómez, José D</p> <p>2014-12-16</p> <p><span class="hlt">Na</span>(+)/Ca(<span class="hlt">2</span>+) exchangers (NCXs) are ubiquitous membrane transporters with a key role in Ca(<span class="hlt">2</span>+) homeostasis and signaling. NCXs mediate the bidirectional translocation of either <span class="hlt">Na</span>(+) or Ca(<span class="hlt">2</span>+), and thus can catalyze uphill Ca(<span class="hlt">2</span>+) transport driven by a <span class="hlt">Na</span>(+) gradient, or vice versa. In a major breakthrough, a prokaryotic NCX homolog (NCX_Mj) was recently isolated and its crystal structure determined at atomic resolution. The structure revealed an intriguing architecture consisting of two inverted-topology repeats, each comprising five transmembrane helices. These repeats adopt asymmetric conformations, yielding an outward-facing occluded state. The crystal structure also revealed four putative ion-binding sites, but the occupancy and specificity thereof could not be conclusively established. Here, we use molecular-dynamics simulations and free-energy calculations to identify the ion configuration that best corresponds to the crystallographic data and that is also thermodynamically optimal. In this most probable configuration, three <span class="hlt">Na</span>(+) ions occupy the so-called Sext, SCa, and Sint sites, whereas the Smid site is occupied by one water molecule and one H(+), which protonates an adjacent aspartate side chain (D240). Experimental measurements of <span class="hlt">Na</span>(+)/Ca(<span class="hlt">2</span>+) and Ca(<span class="hlt">2</span>+)/Ca(<span class="hlt">2</span>+) exchange by wild-type and mutagenized NCX_Mj confirm that transport of both <span class="hlt">Na</span>(+) and Ca(<span class="hlt">2</span>+) requires protonation of D240, and that this side chain does not coordinate either ion at Smid. These results imply that the ion exchange stoichiometry of NCX_Mj is 3:1 and that translocation of <span class="hlt">Na</span>(+) across the membrane is electrogenic, whereas transport of Ca(<span class="hlt">2</span>+) is not. Altogether, these findings provide the basis for further experimental and computational studies of the conformational mechanism of this exchanger.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22658230-high-pressure-structural-investigation-alluaudites-na-sub-fe-sub-po-sub-sub-na-sub-femn-sub-po-sub-sub-system','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22658230-high-pressure-structural-investigation-alluaudites-na-sub-fe-sub-po-sub-sub-na-sub-femn-sub-po-sub-sub-system"><span>High pressure structural investigation on alluaudites <span class="hlt">Na</span>{sub <span class="hlt">2</span>}Fe{sub 3}(PO{sub 4}){sub 3}-<span class="hlt">Na</span>{sub <span class="hlt">2</span>}FeMn{sub <span class="hlt">2</span>}(PO{sub 4}){sub 3} system</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Gao, Jing; Huang, Weifeng; Qin, Shan</p> <p></p> <p>Alluaudites are promising electrochemical materials benefited from the open structure. Structural variations of alluaudites <span class="hlt">Na</span>{sub <span class="hlt">2</span>}M{sub 3}(PO{sub 4}){sub 3} (M{sub 3}=Fe{sub 3}, Fe{sub <span class="hlt">2</span>}Mn and FeMn{sub <span class="hlt">2</span>}) system have been studied by synchrotron radiation X-ray diffraction combined with diamond anvil cell technique up to ~10 GPa at room temperature. No phase transition is observed. The excellent structural stability is mainly due to the flexible framework plus strong covalent P-O bond. Mn{sup <span class="hlt">2</span>+} instead of Fe can be described as <span class="hlt">Na</span>{sup +}+<span class="hlt">2</span>Fe{sup <span class="hlt">2</span>+}→Mn{sup <span class="hlt">2</span>+}+Fe{sup 3+}+□ where □ represents a lattice vacancy. The replacement of Fe with larger Mn{sup <span class="hlt">2</span>+} is equivalentmore » to applying negative chemical pressure to the material. And it causes a more compressible b-axis, lattice expansion, structural compressibility and intensifies the core/electron-electron interactions of Fe. External pressure effect produces anisotropic lattice shrinkage. Structural considerations related to these variations and promising application prospects are discussed. - Graphical abstract: Figure 1 The crystal structure of alluaudites <span class="hlt">Na</span>{sub <span class="hlt">2</span>}M{sub 3}(PO{sub 4}){sub 3} (M{sub 3}=Fe{sub 3}, Fe{sub <span class="hlt">2</span>}Mn and FeMn{sub <span class="hlt">2</span>}) projected along the c-axis. Alluaudites adopt a flexible framework plus strong covalent P-O bond, which contribute to excellent structural stability up to ~10 GPa. Mn{sup <span class="hlt">2</span>+} instead of Fe can be described as <span class="hlt">Na</span>{sup ++}<span class="hlt">2</span>Fe{sup <span class="hlt">2</span>+}→Mn{sup <span class="hlt">2</span>+}+Fe{sup 3+}+□ where □ represents a lattice vacancy, and it is equivalent to applying negative chemical pressure to the host. The substitution causes a more compressible b-axis, lattice expansion, structural compressibility and intensifies the core/electron-electron interactions of Fe.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2228682','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2228682"><span>Transmembrane <span class="hlt">Na</span>+ and Ca<span class="hlt">2</span>+ electrochemical gradients in cardiac muscle and their relationship to force development</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p></p> <p>1982-01-01</p> <p><span class="hlt">Na</span>+- and CA<span class="hlt">2</span>+-sensitive microelectrodes were used to measure intracellular <span class="hlt">Na</span>+ and Ca<span class="hlt">2</span>+ activities (alpha iCa) of sheep ventricular muscle and Purkinje strands to study the interrelationship between <span class="hlt">Na</span>+ and Ca<span class="hlt">2</span>+ electrochemical gradients (delta mu<span class="hlt">Na</span> and delta muCa) under various conditions. In ventricular muscle, alpha i<span class="hlt">Na</span> was 6.4 +/- 1.<span class="hlt">2</span> mM and alpha iCa was 87 +/- 20 nM ([Ca/+] = 272 nM). A graded decrease of external <span class="hlt">Na</span>+ activity (alpha o<span class="hlt">Na</span>) resulted in decrease of alpha i<span class="hlt">Na</span>, and increase of alpha iCa. There was increase of twitch tension in low- alpha o<span class="hlt">Na</span> solutions, and occasional increase of resting tension in 40% alpha o<span class="hlt">Na</span>. Increase of external Ca<span class="hlt">2</span>+ (alpha oCa) resulted in increase of alpha iCa and decrease of alpha i<span class="hlt">Na</span>. Decrease of alpha oCa resulted in decrease of alpha iCa and increase of alpha i<span class="hlt">Na</span>. The apparent resting <span class="hlt">Na</span>-Ca energy ratio (delta muCa/delta mu<span class="hlt">Na</span>) was between <span class="hlt">2</span>.43 and <span class="hlt">2</span>.63. When the membrane potential (Vm) was depolarized by 50 mM K+ in ventricular muscle, Vm depolarized by 50 mV, alpha i<span class="hlt">Na</span> decreased, and alpha iCa increased, with the development of a contracture. The apparent energy coupling ratio did not change with depolarization. 5 x 10(-6) M ouabain induced a large increase in alpha i<span class="hlt">Na</span> ad alpha iCa, accompanied by an increase in twitch and resting tension. Under the conditions we have studied, delta mu<span class="hlt">Na</span> and delta muCa appeared to be coupled and n was nearly constant at <span class="hlt">2</span>.5, as would be expected if the <span class="hlt">Na</span>-Ca exchange system was able to set the steady level of alpha iCa. Tension threshold was about 230 nM alpha iCa. The magnitude of twitch tension was directly related to alpha iCa. PMID:6292328</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19950012911&hterms=Mg+Ca&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3DMg%2BCa','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19950012911&hterms=Mg+Ca&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3DMg%2BCa"><span>Heating during solar nebula formation and Mg isotopic fractionation in precursor grains of <span class="hlt">CAIs</span> and chondrules</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Sasaki, S.; Nagahara, H.; Kitagami, K.; Nakagawa, Y.</p> <p>1994-01-01</p> <p>In some Ca-Al-rich inclusion (<span class="hlt">CAI</span>) grains, mass-dependent isotopic fractionations of Mg, Si, and O are observed and large Mg isotopic fractionation is interpreted to have been produced by cosmochemical processes such as evaporation and condensation. Mass-dependent Mg isotopic fractionation was found in olivine chondrules of Allende meteorites. Presented is an approximate formula for the temperature of the solar nebula that depends on heliocentric distance and the initial gas distribution. Shock heating during solar nebula formation can cause evaporative fractionation within interstellar grains involved in a gas at the inner zone (a less than 3 AU) of the disk. Alternatively collision of late-accreting gas blobs might cause similar heating if Sigma(sub s) and Sigma are large enough. Since the grain size is small, the solid/gas mass ratio is low and solar (low P(sub O<span class="hlt">2</span>)), and the ambient gas pressure is low, this heating event could not produce chondrules themselves. Chondrule formation should proceed around the disk midplane after dust grains would grow and sediment to increase the solid/gas ratio there. The heating source there is uncertain, but transient rapid accretion through the disk could release a large amount of heat, which would be observed as FU Orionis events.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JPS...374...40W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JPS...374...40W"><span>Synthesis of carbon-coated <span class="hlt">Na</span><span class="hlt">2</span>MnPO4F hollow spheres as a potential cathode material for <span class="hlt">Na</span>-ion batteries</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wu, Ling; Hu, Yong; Zhang, Xiaoping; Liu, Jiequn; Zhu, Xing; Zhong, Shengkui</p> <p>2018-01-01</p> <p>Hollow sphere structure <span class="hlt">Na</span><span class="hlt">2</span>MnPO4F/C composite is synthesized through spray drying, following in-situ pyrolytic carbon coating process. XRD results indicate that the well crystallized composite can be successfully synthesized, and no other impurity phases are detected. SEM and TEM results reveal that the <span class="hlt">Na</span><span class="hlt">2</span>MnPO4F/C samples show intact hollow spherical architecture, and the hollow spherical shells with an average thickness of 150 nm-250 nm are composed of nanosized primary particles. Furthermore, the amorphous carbon layer is uniformly coated on the surface of the hollow sphere, and the nanosized <span class="hlt">Na</span><span class="hlt">2</span>MnPO4F particles are well embedded in the carbon networks. Consequently, the hollow sphere structure <span class="hlt">Na</span><span class="hlt">2</span>MnPO4F/C shows enhanced electrochemical performance. Especially, it is the first time that the obvious potential platforms (∼3.6 V) are observed during the charge and discharge process at room temperature.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/6252550-expression-cardiac-sarcolemmal-na-sup-ca-sup-exchange-activity-xenopus-laevis-oocytes','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/6252550-expression-cardiac-sarcolemmal-na-sup-ca-sup-exchange-activity-xenopus-laevis-oocytes"><span>Expression of cardiac sarcolemmal <span class="hlt">Na</span> sup + -Ca sup <span class="hlt">2</span>+ exchange activity in Xenopus laevis oocytes</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>Longoni, S.; Coady, M.J.; Ikeda, T.</p> <p>1988-12-01</p> <p>Injection of Xenopus laevis oocytes with rabbit heart poly(A){sup +}RNA results in expression of <span class="hlt">Na</span>{sup +} inside (<span class="hlt">Na</span>{sub i}{sup +})-dependent Ca{sup <span class="hlt">2</span>+} uptake activity. The activity was measured by first loading the oocytes with <span class="hlt">Na</span>{sup +} using nystatin and then incubating the oocytes in K{sup +} or <span class="hlt">Na</span>{sup +} medium containing {sup 45}Ca. The expressed <span class="hlt">Na</span>{sup +} gradient-dependent Ca{sup <span class="hlt">2</span>+} uptake was five to eight times that observed with water-injected oocytes or with poly(A){sup +}RNA-injected oocytes for which the <span class="hlt">Na</span>{sup +} load step had been omitted. Induced activity was related to the amount of RNA injected and was insensitive tomore » nifedipine. Fractionation of the poly(A){sup +}RNA on a sucrose gradient determined that the active message had a size range between 3 and 8 kb. The properties of the <span class="hlt">Na</span>{sup +} gradient-dependent Ca{sup <span class="hlt">2</span>+} uptake indicated that <span class="hlt">Na</span>{sup +}-Ca{sup <span class="hlt">2</span>+} exchange activity had been expressed in X. laevis oocytes. The result may be useful for cloning and identifying the molecular component responsible for <span class="hlt">Na</span>{sup +}-Ca{sup <span class="hlt">2</span>+} exchange.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JNuM..490..101I','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JNuM..490..101I"><span>Thermal expansion of the nuclear fuel-sodium reaction product <span class="hlt">Na</span>3(U0.84(<span class="hlt">2</span>),<span class="hlt">Na</span>0.16(<span class="hlt">2</span>))O4 - Structural mechanism and comparison with related sodium-metal ternary oxides</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Illy, Marie-Claire; Smith, Anna L.; Wallez, Gilles; Raison, Philippe E.; Caciuffo, Roberto; Konings, Rudy J. M.</p> <p>2017-07-01</p> <p><span class="hlt">Na</span>3.16(<span class="hlt">2</span>)UV,VI0.84(<span class="hlt">2</span>)O4 is obtained from the reaction of sodium with uranium dioxide under oxygen potential conditions typical of a sodium-cooled fast nuclear reactor. In the event of a breach of the steel cladding, it would be the dominant reaction product forming at the rim of the mixed (U,Pu)O<span class="hlt">2</span> fuel pellets. High-temperature X-ray diffraction measurements show that a distortion of the uranium environment in <span class="hlt">Na</span>3.16(<span class="hlt">2</span>)UV,VI0.84(<span class="hlt">2</span>)O4 results in a strongly anisotropic thermal expansion. A comparison with several related sodium metallates Nan-<span class="hlt">2</span>Mn+On-1 - including <span class="hlt">Na</span>3SbO4 and <span class="hlt">Na</span>3TaO4, whose crystal structures are reported for the first time - has allowed us to assess the role played in the lattice expansion by the Mn+ cation radius and the <span class="hlt">Na</span>/M ratio. On this basis, the thermomechanical behavior of the title compound is discussed, along with those of several related double oxides of sodium and actinide elements, surrogate elements, or fission products.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017ApPhL.110j3901Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017ApPhL.110j3901Y"><span>Fast sodium ionic conduction in <span class="hlt">Na</span><span class="hlt">2</span>B10H10-<span class="hlt">Na</span><span class="hlt">2</span>B12H12 pseudo-binary complex hydride and application to a bulk-type all-solid-state battery</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yoshida, Koji; Sato, Toyoto; Unemoto, Atsushi; Matsuo, Motoaki; Ikeshoji, Tamio; Udovic, Terrence J.; Orimo, Shin-ichi</p> <p>2017-03-01</p> <p>In the present work, we developed highly sodium-ion conductive <span class="hlt">Na</span><span class="hlt">2</span>B10H10-<span class="hlt">Na</span><span class="hlt">2</span>B12H12 pseudo-binary complex hydride via mechanically ball-milling admixtures of the pure <span class="hlt">Na</span><span class="hlt">2</span>B10H10 and <span class="hlt">Na</span><span class="hlt">2</span>B12H12 components. Both of these components show a monoclinic phase at room temperature, but ball-milled mixtures partially stabilized highly ion-conductive, disordered cubic phases, whose fraction and favored structural symmetry (body-centered cubic or face-centered cubic) depended on the conditions of mechanical ball-milling and molar ratio of the component compounds. First-principles molecular-dynamics simulations demonstrated that the total energy of the closo-borane mixtures and pure materials is quite close, helping to explain the observed stabilization of the mixed compounds. The ionic conductivity of the closo-borane mixtures appeared to be correlated with the fraction of the body-centered-cubic phase, exhibiting a maximum at a molar ratio of <span class="hlt">Na</span><span class="hlt">2</span>B10H10:<span class="hlt">Na</span><span class="hlt">2</span>B12H12 = 1:3. A conductivity as high as log(σ/S cm-1) = -3.5 was observed for the above ratio at 303 K, being approximately <span class="hlt">2</span>-3 orders of magnitude higher than that of either pure material. A bulk-type all-solid-state sodium-ion battery with a closo-borane-mixture electrolyte, sodium-metal negative-electrode, and TiS<span class="hlt">2</span> positive-electrode demonstrated a high specific capacity, close to the theoretical value of <span class="hlt">Na</span>TiS<span class="hlt">2</span> formation and a stable discharge/charge cycling for at least eleven cycles, with a high discharge capacity retention ratio above 91% from the second cycle.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1332364','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1332364"><span>Sodium ion transport mechanisms in antiperovskite electrolytes <span class="hlt">Na</span> 3OBr and <span class="hlt">Na</span> 4OI <span class="hlt">2</span>: An in Situ neutron diffraction study</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>Zhu, Jinlong; Wang, Yonggang; Li, Shuai</p> <p></p> <p><span class="hlt">Na</span>-rich antiperovskites are recently developed solid electrolytes with enhanced sodium ionic conductivity and show promising functionality as a novel solid electrolyte in an all solid-stat battery. In this work, the sodium ionic transport pathways of the parent compound <span class="hlt">Na</span> 3OBr, as well as the modified layered antiperovskite <span class="hlt">Na</span> 4OI <span class="hlt">2</span>, were studied and compared through temperature dependent neutron diffraction combined with the maximum entropy method. In the cubic <span class="hlt">Na</span> 3OBr antiperovskite, the nuclear density distribution maps at 500 K indicate that sodium ions ho within and among oxygen octahedra, and Br - ions are not involved in the tetragonal Namore » 4OI <span class="hlt">2</span> antiperovskite, <span class="hlt">Na</span> ions, which connect octahedra in the ab plane, have the lowest activation energy barrier. In conclusion, the transport of sodium ions along the c axis is assisted by I - ions.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1332364-sodium-ion-transport-mechanisms-antiperovskite-electrolytes-na3obr-na4oi2-situ-neutron-diffraction-study','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1332364-sodium-ion-transport-mechanisms-antiperovskite-electrolytes-na3obr-na4oi2-situ-neutron-diffraction-study"><span>Sodium ion transport mechanisms in antiperovskite electrolytes <span class="hlt">Na</span> 3OBr and <span class="hlt">Na</span> 4OI <span class="hlt">2</span>: An in Situ neutron diffraction study</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Zhu, Jinlong; Wang, Yonggang; Li, Shuai; ...</p> <p>2016-06-02</p> <p><span class="hlt">Na</span>-rich antiperovskites are recently developed solid electrolytes with enhanced sodium ionic conductivity and show promising functionality as a novel solid electrolyte in an all solid-stat battery. In this work, the sodium ionic transport pathways of the parent compound <span class="hlt">Na</span> 3OBr, as well as the modified layered antiperovskite <span class="hlt">Na</span> 4OI <span class="hlt">2</span>, were studied and compared through temperature dependent neutron diffraction combined with the maximum entropy method. In the cubic <span class="hlt">Na</span> 3OBr antiperovskite, the nuclear density distribution maps at 500 K indicate that sodium ions ho within and among oxygen octahedra, and Br - ions are not involved in the tetragonal Namore » 4OI <span class="hlt">2</span> antiperovskite, <span class="hlt">Na</span> ions, which connect octahedra in the ab plane, have the lowest activation energy barrier. In conclusion, the transport of sodium ions along the c axis is assisted by I - ions.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/1388626-preparation-structure-na2ag5fe3-p2o7-ag-metal-composite-insights-electrochemistry','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/1388626-preparation-structure-na2ag5fe3-p2o7-ag-metal-composite-insights-electrochemistry"><span>Preparation and structure of <span class="hlt">Na</span><span class="hlt">2</span>Ag5Fe3(P<span class="hlt">2</span>O7)4 -Ag metal composite: Insights on electrochemistry</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, Yiman; Marschilok, Amy C.; Takeuchi, Esther S.</p> <p></p> <p>ABSTRACT Ag 7Fe 3(P <span class="hlt">2</span>O 7) 4is a 3D structured material which has been recently studied as a possible cathode material for lithium batteries. Notably, <span class="hlt">Na</span> 7Fe 3(P <span class="hlt">2</span>O 7) 4is reported to be a fast-ion conductor, yet poor electrical conductor. Here, partial replacement of <span class="hlt">Na</span> +for Ag +yielded <span class="hlt">Na</span> <span class="hlt">2</span>Ag 5Fe 3(P <span class="hlt">2</span>O 7) 4pyrophosphate framework where the formation of Ag metal is proposed to increase the intrinsic low electrical conductivity of this polyanion electrode. Specifically, the Ag 5<span class="hlt">Na</span> <span class="hlt">2</span>Fe 3(P <span class="hlt">2</span>O 7) 4-Ag composite is synthesized via chemical reduction of Ag 7Fe 3(P <span class="hlt">2</span>O 7) 4using <span class="hlt">Na</span>BH 4.more » The occupancy of Ag +and <span class="hlt">Na</span> +in each site was determined via Rietveld analysis of the diffraction pattern. Electrochemistry of the Ag 5<span class="hlt">Na</span> <span class="hlt">2</span>Fe 3(P <span class="hlt">2</span>O 7) 4-Ag metal composite was explored with voltammetry and galvanostatic charge/discharge cycling. The Ag 5<span class="hlt">Na</span> <span class="hlt">2</span>Fe 3(P <span class="hlt">2</span>O 7) 4-Ag metal composite electrodes displayed good rate capability assisted by the presence of Ag metal from the chemical reduction and in-situ electrochemical formation of a Ag conductive network.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28581705','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28581705"><span>Characteristics of <span class="hlt">Na</span>NO3-Promoted CdO as a Midtemperature CO<span class="hlt">2</span> Absorbent.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kim, Kang-Yeong; Kwak, Jin-Su; An, Young-In; Oh, Kyung-Ryul; Kwon, Young-Uk</p> <p>2017-06-28</p> <p>In this study, we explored the reaction system CdO(s) + CO <span class="hlt">2</span> (g) ⇄ CdCO 3 (s) as a model system for CO <span class="hlt">2</span> capture agent in the intermediate temperature range of 300-400 °C. While pure CdO does not react with CO <span class="hlt">2</span> at all up to 500 °C, CdO mixed with an appropriate amount of <span class="hlt">Na</span>NO 3 (optimal molar ratio <span class="hlt">Na</span>NO 3 /CdO = 0.14) greatly enhances the conversion of CdO into CdCO 3 up to ∼80% (5.68 mmol/g). These <span class="hlt">Na</span>NO 3 -promoted CdO absorbents can undergo many cycles of absorption and desorption by temperature swing between 300 and 370 °C under a 100% CO <span class="hlt">2</span> condition. Details of how <span class="hlt">Na</span>NO 3 promotes the CO <span class="hlt">2</span> absorption of CdO have been delineated through various techniques using thermogravimetry, coupled with X-ray diffraction and electron microscopy. On the basis of the observed data, we propose a mechanism of CO <span class="hlt">2</span> absorption and desorption of <span class="hlt">Na</span>NO 3 -promoted CdO. The absorption proceeds through a sequence of events of CO <span class="hlt">2</span> adsorption on the CdO surface covered by <span class="hlt">Na</span>NO 3 , dissolution of so-formed CdCO 3 , and precipitation of CdCO 3 particles in the <span class="hlt">Na</span>NO 3 medium. The desorption occurs through the decomposition of CdCO 3 in the dissolved state in the <span class="hlt">Na</span>NO 3 medium where CdO nanoparticles are formed dispersed in the <span class="hlt">Na</span>NO 3 medium. The CdO nanoparticles are aggregated into micrometer-large particles with smooth surfaces and regular shapes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017Nanot..28F5602W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017Nanot..28F5602W"><span><span class="hlt">Na</span>Cl-assisted one-step growth of MoS<span class="hlt">2</span>-WS<span class="hlt">2</span> in-plane heterostructures</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wang, Zhan; Xie, Yong; Wang, Haolin; Wu, Ruixue; Nan, Tang; Zhan, Yongjie; Sun, Jing; Jiang, Teng; Zhao, Ying; Lei, Yimin; Yang, Mei; Wang, Weidong; Zhu, Qing; Ma, Xiaohua; Hao, Yue</p> <p>2017-08-01</p> <p>Transition metal dichalcogenides (TMDs) have attracted considerable interest for exploration of next-generation electronics and optoelectronics in recent years. Fabrication of in-plane lateral heterostructures between TMDs has opened up excellent opportunities for engineering two-dimensional materials. The creation of high quality heterostructures with a facile method is highly desirable but it still remains challenging. In this work, we demonstrate a one-step growth method for the construction of high-quality MoS<span class="hlt">2</span>-WS<span class="hlt">2</span> in-plane heterostructures. The synthesis was carried out using ambient pressure chemical vapor deposition (APCVD) with the assistance of sodium chloride (<span class="hlt">Na</span>Cl). It was found that the addition of <span class="hlt">Na</span>Cl played a key role in lowering the growth temperatures, in which the <span class="hlt">Na</span>-containing precursors could be formed and condensed on the substrates to reduce the energy of the reaction. As a result, the growth regimes of MoS<span class="hlt">2</span> and WS<span class="hlt">2</span> are better matched, leading to the formation of in-plane heterostructures in a single step. The heterostructures were proved to be of high quality with a sharp and clear interface. This newly developed strategy with the assistance of <span class="hlt">Na</span>Cl is promising for synthesizing other TMDs and their heterostructures.</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('http://adsabs.harvard.edu/abs/2017GeCoA.201..185K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017GeCoA.201..185K"><span>Calcium-aluminum-rich inclusions recycled during formation of porphyritic chondrules from CH carbonaceous chondrites</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Krot, Alexander N.; Nagashima, Kazuhide; van Kooten, Elishevah M. M.; Bizzarro, Martin</p> <p>2017-03-01</p> <p>We report on the mineralogy, petrography, and O-isotope compositions of ∼60 Ca, Al-rich inclusions (<span class="hlt">CAIs</span>) incompletely melted during formation of porphyritic chondrules from the CH metal-rich carbonaceous chondrites and Isheyevo (CH/CB). These include (i) relict polymineralic <span class="hlt">CAIs</span> in porphyritic chondrules, (ii) <span class="hlt">CAIs</span> surrounded by chondrule-like igneous rims, (iii) igneous pyroxene-rich and Type C-like <span class="hlt">CAIs</span>, and (iv) plagioclase-rich chondrules with clusters of relict spinel grains. 26Al-26Mg systematics were measured in 10 relict <span class="hlt">CAIs</span> and 11 <span class="hlt">CAI</span>-bearing plagioclase-rich chondrules. Based on the mineralogy, the CH <span class="hlt">CAIs</span> incompletely melted during chondrule formation can be divided into grossite-rich (n = 13), hibonite-rich (n = 11), spinel ± melilite-rich (n = 33; these include plagioclase-rich chondrules with clusters of relict spinel grains) types. Mineralogical observations indicate that these <span class="hlt">CAIs</span> were mixed with different proportions of ferromagnesian silicates and experienced incomplete melting and gas-melt interaction during chondrule formation. These processes resulted in partial or complete destruction of the <span class="hlt">CAI</span> Wark-Lovering rims, replacement of melilite by <span class="hlt">Na</span>-bearing plagioclase, and dissolution and overgrowth of nearly end-member spinel by chromium- and iron-bearing spinel. Only two relict <span class="hlt">CAIs</span> and two <span class="hlt">CAI</span>-bearing chondrules show resolvable excess of radiogenic 26Mg; the inferred initial 26Al/27Al ratios are (1.7 ± 1.3) × 10-6, (3.7 ± 3.1) × 10-7, (1.9 ± 0.9) × 10-6 and (4.9 ± <span class="hlt">2</span>.6) × 10-6. There is a large range of Δ17O among the CH <span class="hlt">CAIs</span> incompletely melted during chondrule formation, from ∼-37‰ to ∼-5‰; the unmelted minerals in individual <span class="hlt">CAIs</span>, however, are isotopically uniform and systematically 16O-enriched relative to the host chondrules and chondrule-like igneous rims, which have Δ17O ranging from ∼-7‰ to ∼+4‰. Most of the CH <span class="hlt">CAIs</span> incompletely melted during chondrule formation are mineralogically and isotopically</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29641197','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29641197"><span>Electronic Spectra of Cs<span class="hlt">2</span><span class="hlt">Na</span>Yb(NO<span class="hlt">2</span>)6: Is There Quantum Cutting?</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Luo, Yuxia; Liu, Zhenyu; Hau, Sam Chun-Kit; Yeung, Yau Yuen; Wong, Ka-Leung; Shiu, Kwok Keung; Chen, Xueyuan; Zhu, Haomiao; Bao, Guochen; Tanner, Peter A</p> <p>2018-05-03</p> <p>The crystal structure and electronic spectra of the T h symmetry hexanitritoytterbate(III) anion have been studied in Cs <span class="hlt">2</span> <span class="hlt">Na</span>Y 0.96 Yb 0.04 (NO <span class="hlt">2</span> ) 6 , which crystallizes in the cubic space group Fm3̅. The emission from Yb 3+ can be excited via the NO <span class="hlt">2</span> - antenna. The latter electronic transition is situated at more than twice the energy of the former, but at room temperature, one photon absorbed at 470 nm in the triplet state produces no more than one photon emitted. Some degree of quantum cutting is observed at 298 K under 420 nm excitation into the singlet state and at 25 K using excitation into either state. The quantum efficiency is ∼10% at 25 K. The energy level scheme of Yb 3+ has been deduced from excitation and emission spectra and calculated by crystal field theory. New improved energy level calculations are also reported for the Cs <span class="hlt">2</span> <span class="hlt">Na</span>Ln(NO <span class="hlt">2</span> ) 6 (Ln = Pr, Eu, Tb) series using the f- Spectra package. The neat crystal Cs <span class="hlt">2</span> <span class="hlt">Na</span>Yb(NO <span class="hlt">2</span> ) 6 has also been studied, but results were unsatisfactory due to sample decomposition, and this chemical instability makes it unsuitable for applications.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27751228','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27751228"><span>[Effect of leptin on expression of calpain-1 and Bcl-<span class="hlt">2</span> and apoptosis in myocardial tissue of neonatal rats after asphyxia].</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wu, Dan-Dan; Wu, Xing-Heng; Zhang, Li-Na</p> <p>2016-10-01</p> <p>To study the effect of leptin on the expression of calcium-activated neutral protease 1 (calpain-1) and B cell lymphoma-<span class="hlt">2</span> (Bcl-<span class="hlt">2</span>) and apoptosis in the myocardial tissue of neonatal rats after asphyxia. A total of 48 neonatal rats were randomly and equally divided into normal control group, asphyxia group, leptin treatment groups, and calpain-1 inhibitor (<span class="hlt">CAI</span>-1) group. The neonatal rat model of asphyxia under normal atmospheric condition was established in all groups except the control group. For the leptin treatment groups, rats received 20, 80, and 160 μg/kg leptin by intraperitoneal injection immediately after model establishment, respectively. For the <span class="hlt">CAI</span>-1 group, rats received 10 mg/kg <span class="hlt">CAI</span>-1 by intraperitoneal injection immediately after model establishment. For all the groups, the myocardial tissue was collected at <span class="hlt">2</span> hours after model establishment. Immunohistochemistry was used to measure the expression of calpain-1 and Bcl-<span class="hlt">2</span>. The TUNEL method was used to evaluate apoptosis of myocardial cells. The expression of calpain-1 and Bcl-<span class="hlt">2</span> and apoptosis index (AI) were significantly higher in the asphyxia group than in the normal control group (P˂0.05). The leptin treatment groups and the <span class="hlt">CAI</span>-1 group had significantly lower expression of calpain-1, significantly lower AI, and significantly higher expression of Bcl-<span class="hlt">2</span> than the asphyxia group (P˂0.05). The <span class="hlt">CAI</span>-1 group had the largest changes in all the indices compared with the asphyxia group. However, there were no significant differences in all indices between the 160 μg/kg leptin treatment group and the <span class="hlt">CAI</span>-1 group. After asphyxia, the expression of calpain-1 was positively correlated with AI, while the expression of Bcl-<span class="hlt">2</span> was negatively correlated with AI and the expression of calpain-1 (P˂0.05). Leptin reduces apoptosis of myocardial cells in asphyxiated neonatal rats by the inhibition of calpain-1 activation and upregulation of Bcl-<span class="hlt">2</span> expression.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.dtic.mil/docs/citations/ADP001715','DTIC-ST'); return false;" href="http://www.dtic.mil/docs/citations/ADP001715"><span>Microplane Model for Fracture Analysis of Concrete Structures</span></a></p> <p><a target="_blank" href="http://www.dtic.mil/">DTIC Science & Technology</a></p> <p></p> <p>1983-05-01</p> <p>Curves for Cot,- Proceadings, Intern. Sympositim <span class="hlt">cai</span> Numevical crete io Tension," Mater ils ,tndI : Str ~tuiztres d.- Un R. /<span class="hlt">na</span>~gr No I., an. o196.9...and 226. ivinsterwalder, K, , "Festigkeit und Verfor- mung van Bet an lniter- 7.oRsIannitgenl, Duo tsatho 8. Kr~ner, h. , Zur <span class="hlt">2</span>’lastischen Verformung</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22591357-structure-microstructure-infrared-studies-ba-sub-na-sub-bi-sub-sub-tio-sub-nanbo-sub-ceramics','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22591357-structure-microstructure-infrared-studies-ba-sub-na-sub-bi-sub-sub-tio-sub-nanbo-sub-ceramics"><span>Structure, microstructure and infrared studies of Ba{sub 0.06}(<span class="hlt">Na</span>{sub 1/<span class="hlt">2</span>}Bi{sub 1/<span class="hlt">2</span>}){sub 0.94}TiO{sub 3}-<span class="hlt">Na</span>NbO{sub 3} ceramics</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>Roy, Sumit K., E-mail: sumit.sxc13@gmail.com; Singh, S. N., E-mail: snsphyru@gmail.com; Prasad, K., E-mail: k.prasad65@gmail.com</p> <p>2016-05-06</p> <p>Lead-free solid solutions (1-x)Ba{sub 0.06}(<span class="hlt">Na</span>{sub 1/<span class="hlt">2</span>}Bi{sub 1/<span class="hlt">2</span>}){sub 0.94}TiO{sub 3}-x<span class="hlt">Na</span>NbO{sub 3} (0 ≤ x ≤ 1.0) were prepared by conventional ceramic fabrication technique. X-ray diffraction and Rietveld refinement analyses of these ceramics were carried out using X’Pert HighScore Plus software to determine the crystal symmetry, space group and unit cell dimensions. Rietveld refinement revealed that <span class="hlt">Na</span>NbO{sub 3} with orthorhombic structure was completely diffused into Ba{sub 0.06}(<span class="hlt">Na</span>{sub 1/<span class="hlt">2</span>}Bi{sub 1/<span class="hlt">2</span>}){sub 0.94}TiO{sub 3} lattice having the rhombohedral-tetragonal symmetry. EDS and SEM studies were carried out in order to evaluate the quality and purity of the compounds. SEM images showed a change in grain shapemore » with the increase of <span class="hlt">Na</span>NbO{sub 3} content. FTIR spectra confirmed the formation of solid solution.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4273333','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4273333"><span>Sodium recognition by the <span class="hlt">Na</span>+/Ca<span class="hlt">2</span>+ exchanger in the outward-facing conformation</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Marinelli, Fabrizio; Almagor, Lior; Hiller, Reuben; Giladi, Moshe; Khananshvili, Daniel; Faraldo-Gómez, José D.</p> <p>2014-01-01</p> <p><span class="hlt">Na</span>+/Ca<span class="hlt">2</span>+ exchangers (NCXs) are ubiquitous membrane transporters with a key role in Ca<span class="hlt">2</span>+ homeostasis and signaling. NCXs mediate the bidirectional translocation of either <span class="hlt">Na</span>+ or Ca<span class="hlt">2</span>+, and thus can catalyze uphill Ca<span class="hlt">2</span>+ transport driven by a <span class="hlt">Na</span>+ gradient, or vice versa. In a major breakthrough, a prokaryotic NCX homolog (NCX_Mj) was recently isolated and its crystal structure determined at atomic resolution. The structure revealed an intriguing architecture consisting of two inverted-topology repeats, each comprising five transmembrane helices. These repeats adopt asymmetric conformations, yielding an outward-facing occluded state. The crystal structure also revealed four putative ion-binding sites, but the occupancy and specificity thereof could not be conclusively established. Here, we use molecular-dynamics simulations and free-energy calculations to identify the ion configuration that best corresponds to the crystallographic data and that is also thermodynamically optimal. In this most probable configuration, three <span class="hlt">Na</span>+ ions occupy the so-called Sext, SCa, and Sint sites, whereas the Smid site is occupied by one water molecule and one H+, which protonates an adjacent aspartate side chain (D240). Experimental measurements of <span class="hlt">Na</span>+/Ca<span class="hlt">2</span>+ and Ca<span class="hlt">2</span>+/Ca<span class="hlt">2</span>+ exchange by wild-type and mutagenized NCX_Mj confirm that transport of both <span class="hlt">Na</span>+ and Ca<span class="hlt">2</span>+ requires protonation of D240, and that this side chain does not coordinate either ion at Smid. These results imply that the ion exchange stoichiometry of NCX_Mj is 3:1 and that translocation of <span class="hlt">Na</span>+ across the membrane is electrogenic, whereas transport of Ca<span class="hlt">2</span>+ is not. Altogether, these findings provide the basis for further experimental and computational studies of the conformational mechanism of this exchanger. PMID:25468964</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28635039','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28635039"><span>Rational Design of <span class="hlt">Na</span>(Li1/3 Mn<span class="hlt">2</span>/3 )O<span class="hlt">2</span> Operated by Anionic Redox Reactions for Advanced Sodium-Ion Batteries.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kim, Duho; Cho, Maenghyo; Cho, Kyeongjae</p> <p>2017-09-01</p> <p>In an effort to develop high-energy-density cathodes for sodium-ion batteries (SIBs), low-cost, high capacity <span class="hlt">Na</span>(Li 1/3 Mn <span class="hlt">2</span>/3 )O <span class="hlt">2</span> is discovered, which utilizes the labile O <span class="hlt">2</span>p-electron for charge compensation during the intercalation process, inspired by Li <span class="hlt">2</span> MnO 3 redox reactions. <span class="hlt">Na</span>(Li 1/3 Mn <span class="hlt">2</span>/3 )O <span class="hlt">2</span> is systematically designed by first-principles calculations considering the Li/<span class="hlt">Na</span> mixing enthalpy based on the site preference of <span class="hlt">Na</span> in the Li sites of Li <span class="hlt">2</span> MnO 3 . Using the anionic redox reaction (O <span class="hlt">2</span>- /O - ), this Mn-oxide is predicted to show high redox potentials (≈4.<span class="hlt">2</span> V vs <span class="hlt">Na/Na</span> + ) with high charge capacity (190 mAh g -1 ). Predicted cathode performance is validated by experimental synthesis, characterization, and cyclic performance studies. Through a fundamental understanding of the redox reaction mechanism in Li <span class="hlt">2</span> MnO 3 , <span class="hlt">Na</span>(Li 1/3 Mn <span class="hlt">2</span>/3 )O <span class="hlt">2</span> is designed as an example of a new class of promising cathode materials, <span class="hlt">Na</span>(Li 1/3 M <span class="hlt">2</span>/3 )O <span class="hlt">2</span> (M: transition metals featuring stabilized M 4+ ), for further advances in SIBs. © 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/15090193','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/15090193"><span>Jak<span class="hlt">2</span> and Ca<span class="hlt">2</span>+/calmodulin are key intermediates for bradykinin B<span class="hlt">2</span> receptor-mediated activation of <span class="hlt">Na</span>+/H+ exchange in KNRK and CHO cells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Lefler, David; Mukhin, Yurii V; Pettus, Tobiah; Leeb-Lundberg, L M Fredrik; Garnovskaya, Maria N; Raymond, John R</p> <p>2003-04-01</p> <p><span class="hlt">Na</span>(+)/H(+) exchangers are ubiquitous in mammalian cells, carrying out key functions, such as cell volume defense, acid-base homeostasis, and regulation of the cytoskeleton. We used two screening technologies (FLIPR and microphysiometry) to characterize the signal transduction pathway used by the bradykinin B(<span class="hlt">2</span>) receptor to activate <span class="hlt">Na</span>(+)/H(+) exchange in two cell lines, KNRK and CHO. In both cell types, B(<span class="hlt">2</span>) receptor activation resulted in rapid increases in the rate of proton extrusion that were sodium-dependent and could be blocked by the <span class="hlt">Na</span>(+)/H(+) exchange inhibitors EIPA and MIA or by replacing extracellular sodium with TMA. Activation of <span class="hlt">Na</span>(+)/H(+) exchange by bradykinin was concentration-dependent and could be blocked by the selective B(<span class="hlt">2</span>) receptor antagonist HOE140, but not by the B(1) receptor antagonist des-Arg10-HOE140. Inhibitors of Jak<span class="hlt">2</span> tyrosine kinase (genistein and AG490) and of CAM (W-7 and calmidazolium) attenuated bradykinin-induced activation of <span class="hlt">Na</span>(+)/H(+) exchange. Bradykinin induced formation of a complex between CAM and Jak<span class="hlt">2</span>, supporting a regulatory role for Jak<span class="hlt">2</span> and CAM in the activation of <span class="hlt">Na</span>(+)/H(+) exchange in KNRK and CHO cells. We propose that this pathway (B(<span class="hlt">2</span>) receptor --> Jak<span class="hlt">2</span> --> CAM --> <span class="hlt">Na</span>(+)/H(+) exchanger) is a fundamental regulator of <span class="hlt">Na</span>(+)/H(+) exchange activity.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JPS...372..270Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JPS...372..270Z"><span><span class="hlt">Na</span>3.4Zr1.8Mg0.<span class="hlt">2</span>Si<span class="hlt">2</span>PO12 filled poly(ethylene oxide)/<span class="hlt">Na</span>(CF3SO<span class="hlt">2)2</span>N as flexible composite polymer electrolyte for solid-state sodium batteries</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zhang, Zhizhen; Xu, Kaiqi; Rong, Xiaohui; Hu, Yong-Sheng; Li, Hong; Huang, Xuejie; Chen, Liquan</p> <p>2017-12-01</p> <p>Solid electrolytes with high ionic conductivity and excellent electrochemical stability are of prime significance to enable the application of solid-state batteries in energy storage and conversion. In this study, solid composite polymer electrolytes (CPEs) based on sodium bis(trifluorosulfonyl) imide (<span class="hlt">Na</span>TFSI) and poly (ethylene oxide) (PEO) incorporated with active ceramic filler (NASICON) are reported for the first time. With the addition of NASICON fillers, the thermal stability and electrochemical stability of the CPEs are improved. A high conductivity of <span class="hlt">2</span>.8 mS/cm (at 80 °C) is readily achieved when the content of the NASICON filler in the composite polymer reaches 50 wt%. Furthermore, <span class="hlt">Na</span>3V<span class="hlt">2</span>(PO4)3/CPE/<span class="hlt">Na</span> solid-state batteries using this composite electrolyte display good rate and excellent cycle performance.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/6116010-crystal-structure-true-nasicon-na-sub-zr-sub-si-sub-po-sub','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/6116010-crystal-structure-true-nasicon-na-sub-zr-sub-si-sub-po-sub"><span>Crystal structure of the true Nasicon: <span class="hlt">Na</span>/sub 3/Zr/sub <span class="hlt">2</span>/Si/sub <span class="hlt">2</span>/PO/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>Boilot, J.P.; Collin, G.; Colomban, P.</p> <p></p> <p>For the first time, the results of single crystal determination of the true Nasicon are given. The structure refinement yielded the following composition: <span class="hlt">Na</span>/sub 3.09(8)/Zr/sub <span class="hlt">2</span>.01(1)/P/sub 0.91/Si/sub <span class="hlt">2</span>.09/O/sub 12/. Evidence of the total occupancy of the Zr octahedron is found, displaying that only the Si/P non-stoichiometry mechanism is present in the Nasicon crystal. For the two temperatures which have been investigated (R.T. and 623K), the structures are very close to that of the Nasicon analog: <span class="hlt">Na</span>/sub 3/Sc/sub <span class="hlt">2</span>/P/sub 3/O/sub 12/. However the Si/P substitution prevents the sodium long range ordering even in the monoclinic low temperature phase and therefore themore » cross over to the rhombohedral symmetry only involves very small atomic displacements. For both structures, a new sodium position (mid-<span class="hlt">Na</span>) is displayed in the conduction channel, intermediate between the usual <span class="hlt">Na</span>(1) and <span class="hlt">Na</span>(<span class="hlt">2</span>) sites.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/14717442','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/14717442"><span>Different blocking effects of HgCl<span class="hlt">2</span> and <span class="hlt">Na</span>Cl on aquaporins of pepper plants.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Martínez-Ballesta, M Carmen; Diaz, Rafael; Martínez, Vicente; Carvajal, Micaela</p> <p>2003-12-01</p> <p>In this study we have compared the short-term effects of both <span class="hlt">Na</span>Cl and HgCl<span class="hlt">2</span> on aquaporins of Capsicum annuum L. plants, in order to determine whether or not they are similar. Stomatal conductance, turgor, root hydraulic conductance and water status were measured after 0.5, <span class="hlt">2</span>, 4 and 6 h of <span class="hlt">Na</span>Cl (60 mmol/L) or HgCl<span class="hlt">2</span> (50 micromol/L) treatment. When 60 mmol/L <span class="hlt">Na</span>Cl was added to the nutrient solution, a large decrease in stomatal conductance was observed after <span class="hlt">2</span> h. However, when HgCl<span class="hlt">2</span> (50 micromol/L) was added, the decrease occurred after 4 h. The number of open stomata closed was always lower in plants treated with HgCl<span class="hlt">2</span> than in plants treated with <span class="hlt">Na</span>Cl. The water content of the Hg(<span class="hlt">2</span>+)-treated plants was decreased, compared with controls and <span class="hlt">Na</span>Cl-treated. The root hydraulic conductance decreased after HgCl<span class="hlt">2</span> and <span class="hlt">Na</span>Cl treatment plants. Turgor of leaf epidermal cells was greatly reduced in plants treated with HgCl<span class="hlt">2</span>, but remained constant in the <span class="hlt">Na</span>Cl treatment, compared with control plants. The fact that the stomatal conductance was reduced more rapidly after <span class="hlt">Na</span>Cl addition, followed by the stomatal closure, and that both water content and turgor did not differ from the control suggests that in <span class="hlt">Na</span>Cl-treated plants there must be a signal moving from root to shoot. Therefore, the control of plant homeostasis through a combined regulation of root and stomatal exchanges may be dependent on aquaporin regulation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24071912','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24071912"><span>Hydrogen-fluorine exchange in <span class="hlt">Na</span>BH4-<span class="hlt">Na</span>BF4.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Rude, L H; Filsø, U; D'Anna, V; Spyratou, A; Richter, B; Hino, S; Zavorotynska, O; Baricco, M; Sørby, M H; Hauback, B C; Hagemann, H; Besenbacher, F; Skibsted, J; Jensen, T R</p> <p>2013-11-07</p> <p>Hydrogen-fluorine exchange in the <span class="hlt">Na</span>BH4-<span class="hlt">Na</span>BF4 system is investigated using a range of experimental methods combined with DFT calculations and a possible mechanism for the reactions is proposed. Fluorine substitution is observed using in situ synchrotron radiation powder X-ray diffraction (SR-PXD) as a new Rock salt type compound with idealized composition <span class="hlt">Na</span>BF<span class="hlt">2</span>H<span class="hlt">2</span> in the temperature range T = 200 to 215 °C. Combined use of solid-state (19)F MAS NMR, FT-IR and DFT calculations supports the formation of a BF<span class="hlt">2</span>H<span class="hlt">2</span>(-) complex ion, reproducing the observation of a (19)F chemical shift at -144.<span class="hlt">2</span> ppm, which is different from that of <span class="hlt">Na</span>BF4 at -159.<span class="hlt">2</span> ppm, along with the new absorption bands observed in the IR spectra. After further heating, the fluorine substituted compound becomes X-ray amorphous and decomposes to <span class="hlt">Na</span>F at ~310 °C. This work shows that fluorine-substituted borohydrides tend to decompose to more stable compounds, e.g. <span class="hlt">Na</span>F and BF3 or amorphous products such as closo-boranes, e.g. <span class="hlt">Na</span><span class="hlt">2</span>B12H12. The <span class="hlt">Na</span>BH4-<span class="hlt">Na</span>BF4 composite decomposes at lower temperatures (300 °C) compared to <span class="hlt">Na</span>BH4 (476 °C), as observed by thermogravimetric analysis. <span class="hlt">Na</span>BH4-<span class="hlt">Na</span>BF4 (1:0.5) preserves 30% of the hydrogen storage capacity after three hydrogen release and uptake cycles compared to 8% for <span class="hlt">Na</span>BH4 as measured using Sievert's method under identical conditions, but more than 50% using prolonged hydrogen absorption time. The reversible hydrogen storage capacity tends to decrease possibly due to the formation of <span class="hlt">Na</span>F and <span class="hlt">Na</span><span class="hlt">2</span>B12H12. On the other hand, the additive sodium fluoride appears to facilitate hydrogen uptake, prevent foaming, phase segregation and loss of material from the sample container for samples of <span class="hlt">Na</span>BH4-<span class="hlt">Na</span>F.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010ApSS..257...25K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010ApSS..257...25K"><span>On the temperature dependence of <span class="hlt">Na</span> migration in thin SiO <span class="hlt">2</span> films during ToF-SIMS O <span class="hlt">2</span>+ depth profiling</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Krivec, Stefan; Detzel, Thomas; Buchmayr, Michael; Hutter, Herbert</p> <p>2010-10-01</p> <p>The detection of <span class="hlt">Na</span> in insulating samples by means of time of flight-secondary ion mass spectrometry (ToF-SIMS) depth profiling has always been a challenge. In particular the use of O <span class="hlt">2</span>+ as sputter species causes a severe artifact in the <span class="hlt">Na</span> depth distribution due to <span class="hlt">Na</span> migration under the influence of an internal electrical filed. In this paper we address the influence of the sample temperature on this artifact. It is shown that the transport of <span class="hlt">Na</span> is a dynamic process in concordance with the proceeding sputter front. Low temperatures mitigated the migration process by reducing the <span class="hlt">Na</span> mobility in the target. In the course of this work two sample types have been investigated: (i) A <span class="hlt">Na</span> doped PMMA layer, deposited on a thin SiO <span class="hlt">2</span> film. Here, the incorporation behavior of <span class="hlt">Na</span> into SiO <span class="hlt">2</span> during depth profiling is demonstrated. (ii) <span class="hlt">Na</span> implanted into a thin SiO <span class="hlt">2</span> film. By this sample type the migration behavior could be examined when defects, originating from the implantation process, are present in the SiO <span class="hlt">2</span> target. In addition, we propose an approach for the evaluation of an implanted <span class="hlt">Na</span> profile, which is unaffected by the migration process.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016MinPe.110..905K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016MinPe.110..905K"><span>On the existence of a high-temperature polymorph of <span class="hlt">Na</span><span class="hlt">2</span>Ca6Si4O15—implications for the phase equilibria in the system <span class="hlt">Na</span><span class="hlt">2</span>O-CaO-SiO<span class="hlt">2</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kahlenberg, Volker; Maier, Matthias</p> <p>2016-12-01</p> <p>Singe crystals of a new high-temperature polymorph of <span class="hlt">Na</span><span class="hlt">2</span>Ca6Si4O15 have been obtained from solid state reactions performed at 1300 °C. The basic crystallographic data of this so-called β-phase at ambient conditions are as follows: space group P1 c1, a = 9.0112(5) Å, b = 7.3171(5) Å, c = 10.9723(6) Å, β = 107.720(14)°, V = 689.14(7) Å3, Z = <span class="hlt">2</span>. The crystals showed twinning by reticular merohedry (mimicking an orthorhombic C-centred unit cell) which was accounted for during data processing and structure solution. Structure determination was accomplished by direct methods. Least-squares refinements resulted in a residual of R(|F|) = 0.043 for 5811 observed reflections with I > <span class="hlt">2</span>σ(I). From a structural point of view β-<span class="hlt">Na</span><span class="hlt">2</span>Ca6Si4O15 can be attributed to the group of mixed-anion silicates containing [Si<span class="hlt">2</span>O7]-dimers as well as isolated [SiO4]-tetrahedra in the ratio 1:<span class="hlt">2</span>, i.e. more precisely the formula can be written as <span class="hlt">Na</span><span class="hlt">2</span>Ca6[SiO4]<span class="hlt">2</span>[Si<span class="hlt">2</span>O7]. The tetrahedral groups are arranged in layers parallel to (100). Sodium and calcium cations are located between the silicate anions for charge compensation and are coordinated by six to eight nearest oxygen ligands. Alternatively, the structure can be described as a mixed tetrahedral-octahedral framework based on kröhnkite-type [Ca(SiO4)<span class="hlt">2</span>O<span class="hlt">2</span>]-chains in which the CaO6-octahedra are corner-linked to bridging SiO4-tetrahedra. The infinite chains are running parallel to [001] and are concentrated in layers parallel to (010). Adjacent layers are shifted relative to each other by an amount of +δ or -δ along a*. Consequently, a …ABABAB… stacking sequence is created. A detailed comparison with related structures such as α-<span class="hlt">Na</span><span class="hlt">2</span>Ca6Si4O15 and other A<span class="hlt">2</span>B6Si4O15 representatives including topological as well as group theoretical aspects is presented. There are strong indications that monoclinic <span class="hlt">Na</span><span class="hlt">2</span>Ca3Si<span class="hlt">2</span>O8 mentioned in earlier studies is actually misinterpreted β-<span class="hlt">Na</span><span class="hlt">2</span>Ca6Si4O15. In addition to the detailed crystallographic analysis of</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/7008315-structure-cristalline-de-naliyb-sub-sub-composes-isotypes','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/7008315-structure-cristalline-de-naliyb-sub-sub-composes-isotypes"><span>Structure cristalline de <span class="hlt">Na</span>LiYb/sub <span class="hlt">2</span>/F/sub 8/: composes isotypes</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>Dib, A.; Gorius, M.F.; Aleonard, S.</p> <p>1986-11-15</p> <p><span class="hlt">Na</span>LiYb/sub <span class="hlt">2</span>/F/sub 8/ crystallizes in the monoclinic system, space group C<span class="hlt">2</span>/c, with a = 10.3516(9), b = 8.2069(9), c = 6,9674(7) A, ..beta.. = 90/sup 0/, Z = 4. The crystal structure has been solved from single crystal diffractometer measurements (AgK..cap alpha.. radiation) using Patterson and Fourier syntheses and refined by a least-squares method. The final R value is 0.021 for 1756 independent observed reflections. Two YbF/sub 8/ polyhedra share one of their edges to form Yb/sub <span class="hlt">2</span>/F/sub 14/ groups which are three-dimensionally linked and create cavities in which <span class="hlt">Na</span> and Li are located. Yb/sub <span class="hlt">2</span>/F/sub 14/ groups share twomore » of their edges with these of one <span class="hlt">Na</span>F/sub 7/ (or <span class="hlt">Na</span>F/sub 9/ polyhedron to form Y/sub <span class="hlt">2</span>/<span class="hlt">Na</span>F/sub 20/ blocks. As well, the structure may be described by the packing of planes formed by these two-dimensionally linked blocks. The description of this structure is compared with that ascribed to <span class="hlt">Na</span>LiY/sub <span class="hlt">2</span>/F/sub 8/, which has been described in the monoclinic system, space group P<span class="hlt">2</span>/sub 1//m, with a' = (a + b)/<span class="hlt">2</span>, b' = c,c' = (b - a)/<span class="hlt">2</span>, Z = <span class="hlt">2</span>. It is shown that positions of the atoms confer pseudoorthorhombic symmetry on the cell. Lattice parameters of isotypic compounds are given in the pseudoorthorhombic lattice and they are compared with those recently published with the structure described in the monoclinic system, space group P<span class="hlt">2</span>/sub 1//m.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70015141','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70015141"><span>Critical behavior of dilute <span class="hlt">Na</span>Cl in H<span class="hlt">2</span>O</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Pitzer, Kenneth S.; Bischoff, J.L.; Rosenbauer, R.J.</p> <p>1987-01-01</p> <p>The compositions of the saturated vapor and liquid phases are measured for the system <span class="hlt">Na</span>Cl-H<span class="hlt">2</span>O at 380??C, which is close to the critical point of pure water. The shape of the phase equilibrium curve is classical, which confirms a conclusion reached earlier on the basis of less accurate data. This implies that the long-range forces introduced by the <span class="hlt">Na</span>Cl suppress the non-classical effects present in pure H<span class="hlt">2</span>O. An empirical equation of a classical type fits these data. ?? 1987.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://eric.ed.gov/?q=dolphins&pg=6&id=ED287161','ERIC'); return false;" href="https://eric.ed.gov/?q=dolphins&pg=6&id=ED287161"><span>Cost-Benefit Analysis for ECIA Chapter 1 and State DPPF Programs Comparing Groups Receiving Regular Program Instruction and Groups Receiving Computer Assisted Instruction/Computer Management System (<span class="hlt">CAI</span>/CMS). 1986-87.</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>Chamberlain, Ed</p> <p></p> <p>A cost benefit study was conducted to determine the effectiveness of a computer assisted instruction/computer management system (<span class="hlt">CAI</span>/CMS) as an alternative to conventional methods of teaching reading within Chapter 1 and DPPF funded programs of the Columbus (Ohio) Public Schools. The Chapter 1 funded Compensatory Language Experiences and Reading…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/1212235-batteries-advanced-na-fecl2-zebra-battery-stationary-energy-storage-application','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/1212235-batteries-advanced-na-fecl2-zebra-battery-stationary-energy-storage-application"><span>Batteries: An Advanced <span class="hlt">Na</span>-FeCl<span class="hlt">2</span> ZEBRA Battery for Stationary Energy Storage Application</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, Guosheng; Lu, Xiaochuan; Kim, Jin Yong</p> <p>2015-06-17</p> <p>Sodium-metal chloride batteries, ZEBRA, are considered as one of the most important electrochemical devices for stationary energy storage applications because of its advantages of good cycle life, safety, and reliability. However, sodium-nickel chloride (<span class="hlt">Na</span>-NiCl<span class="hlt">2</span>) batteries, the most promising redox chemistry in ZEBRA batteries, still face great challenges for the practical application due to its inevitable feature of using Ni cathode (high materials cost). In this work, a novel intermediate-temperature sodium-iron chloride (<span class="hlt">Na</span>-FeCl<span class="hlt">2</span>) battery using a molten sodium anode and Fe cathode is proposed and demonstrated. The first use of unique sulfur-based additives in Fe cathode enables <span class="hlt">Na</span>-FeCl<span class="hlt">2</span> batteries can bemore » assembled in the discharged state and operated at intermediate-temperature (<200°C). The results in this work demonstrate that intermediate-temperature <span class="hlt">Na</span>-FeCl<span class="hlt">2</span> battery technology could be a propitious solution for ZEBRA battery technologies by replacing the traditional <span class="hlt">Na</span>-NiCl<span class="hlt">2</span> chemistry.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1980GeCoA..44.1029P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1980GeCoA..44.1029P"><span>Mineral-solution equilibria—III. The system <span class="hlt">Na</span> <span class="hlt">2</span>OAl <span class="hlt">2</span>O 3SiO <span class="hlt">2</span>H <span class="hlt">2</span>OHCl</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Popp, Robert K.; Frantz, John D.</p> <p>1980-07-01</p> <p>Chemical equilibrium between sodium-aluminum silicate minerals and chloride bearing fluid has been experimentally determined in the range 500-700°C at 1 kbar, using rapid-quench hydrothermal methods and two modifications of the Ag + AgCl acid buffer technique. The temperature dependence of the thermodynamic equilibrium constant ( K) for the reaction <span class="hlt">Na</span>AlSi 3O 8 + HCl o = <span class="hlt">Na</span>Cl o + 1/<span class="hlt">2</span>Al <span class="hlt">2</span>SiO 5, + 5/<span class="hlt">2</span>SiO <span class="hlt">2</span> + 1/<span class="hlt">2</span>H <span class="hlt">2</span>O Albite Andalusite Qtz. K = (a <span class="hlt">Na</span>Cl o) /(a H <span class="hlt">2</span>O ) 1/<span class="hlt">2</span>/(a HCl o) can be described by the following equation: log k = -4.437 + 5205.6/ T( K) The data from this study are consistent with experimental results reported by MONTOYA and HEMLEY (1975) for lower temperature equilibria defined by the assemblages albite + paragonite + quartz + fluid and paragonite + andalusite + quartz + fluid. Values of the equilibrium constants for the above reactions were used to estimate the difference in Gibbs free energy of formation between <span class="hlt">Na</span>Cl o and HCl o in the range 400-700°C and 1-<span class="hlt">2</span> kbar. Similar calculations using data from phase equilibrium studies reported in the literature were made to determine the difference in Gibbs free energy of formation between KCl o and HCl o. These data permit modelling of the chemical interaction between muscovite + kspar + paragonite + albite + quartz assemblages and chloride-bearing hydrothermal fluids.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22584113-syntheses-structural-characterization-non-centrosymmetric-na-sub-sub-sub-ga-si-ge-sn-zn-cd-sulfides','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22584113-syntheses-structural-characterization-non-centrosymmetric-na-sub-sub-sub-ga-si-ge-sn-zn-cd-sulfides"><span>Syntheses and structural characterization of non-centrosymmetric <span class="hlt">Na</span>{sub <span class="hlt">2</span>}M{sub <span class="hlt">2</span>}M'S{sub 6} (M, M′=Ga, In, Si, Ge, Sn, Zn, Cd) sulfides</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>Yohannan, Jinu P.; Vidyasagar, Kanamaluru, E-mail: kvsagar@iitm.ac.in</p> <p></p> <p>Seven new non-centrosymmetric <span class="hlt">Na</span>{sub <span class="hlt">2</span>}M{sub <span class="hlt">2</span>}M’S{sub 6} sulfides, namely, <span class="hlt">Na</span>{sub <span class="hlt">2</span>}Sn{sub <span class="hlt">2</span>}ZnS{sub 6}(1){sub ,} <span class="hlt">Na</span>{sub <span class="hlt">2</span>}Ga{sub <span class="hlt">2</span>}GeS{sub 6}(<span class="hlt">2</span>), <span class="hlt">Na</span>{sub <span class="hlt">2</span>}Ga{sub <span class="hlt">2</span>}SnS{sub 6}(3-α), <span class="hlt">Na</span>{sub <span class="hlt">2</span>}Ga{sub <span class="hlt">2</span>}SnS{sub 6}(3-β){sub ,} <span class="hlt">Na</span>{sub <span class="hlt">2</span>}Ge{sub <span class="hlt">2</span>}ZnS{sub 6}(4){sub ,} <span class="hlt">Na</span>{sub <span class="hlt">2</span>}Ge{sub <span class="hlt">2</span>}CdS{sub 6}(5){sub ,} <span class="hlt">Na</span>{sub <span class="hlt">2</span>}In{sub <span class="hlt">2</span>}SiS{sub 6}(6) and <span class="hlt">Na</span>{sub <span class="hlt">2</span>}In{sub <span class="hlt">2</span>}GeS{sub 6}(7), were synthesized by high temperature solid state reactions and structurally characterized by single crystal X-ray diffraction. They crystallize in non-centrosymmetric Fdd<span class="hlt">2</span> and Cc space groups and their three-dimensional [M{sub <span class="hlt">2</span>}M′S{sub 6}]{sup <span class="hlt">2</span>-}framework structures consist of MS{sub 4} and M′S{sub 4} tetrahedra corner-connected to one another in either orderly or disordered fashion. Sodium ions residemore » in the tunnels of the anionic framework. Compounds 1, <span class="hlt">2</span> and 3-α have the structure of known Li{sub <span class="hlt">2</span>}Ga{sub <span class="hlt">2</span>}GeS{sub 6}, whereas compounds 6 and 7 are isostructural with known Li{sub <span class="hlt">2</span>}In{sub <span class="hlt">2</span>}GeS{sub 6} compound. Isostructural compounds 4 and 5 represent a new structural variant. Compounds 3-α and its new monoclinic structural variant 3-β have disordered structural framework. All of them are wide band gap semiconductors. <span class="hlt">Na</span>{sub <span class="hlt">2</span>}Ga{sub <span class="hlt">2</span>}GeS{sub 6}(<span class="hlt">2</span>), <span class="hlt">Na</span>{sub <span class="hlt">2</span>}Ga{sub <span class="hlt">2</span>}SnS{sub 6}(3), <span class="hlt">Na</span>{sub <span class="hlt">2</span>}Ge{sub <span class="hlt">2</span>}ZnS{sub 6}(4) and <span class="hlt">Na</span>{sub <span class="hlt">2</span>}In{sub <span class="hlt">2</span>}GeS{sub 6}(7) compounds are found to be second-harmonic generation (SHG) active. Compounds 1, <span class="hlt">2</span> and 3-α melt congruently. - Graphical abstract: <span class="hlt">Na</span>{sub <span class="hlt">2</span>}Ga{sub <span class="hlt">2</span>}GeS{sub 6}, <span class="hlt">Na</span>{sub <span class="hlt">2</span>}Ga{sub <span class="hlt">2</span>}SnS{sub 6}, <span class="hlt">Na</span>{sub <span class="hlt">2</span>}Ge{sub <span class="hlt">2</span>}ZnS{sub 6}, <span class="hlt">Na</span>{sub <span class="hlt">2</span>}In{sub <span class="hlt">2</span>}GeS{sub 6}, <span class="hlt">Na</span>{sub <span class="hlt">2</span>}Sn{sub <span class="hlt">2</span>}ZnS{sub 6}, <span class="hlt">Na</span>{sub <span class="hlt">2</span>}Ge{sub <span class="hlt">2</span>}CdS{sub 6} and <span class="hlt">Na</span>{sub <span class="hlt">2</span>}In{sub <span class="hlt">2</span>}SiS{sub 6} have non-centrosymmetric structures and the first four compounds are SHG active. Display Omitted - Highlights: • Seven new <span class="hlt">Na</span>{sub <span class="hlt">2</span>}M{sub <span class="hlt">2</span>}M′S{sub 6} compounds with non-centrosymmetric structures were synthesized. • They are wide band gap semiconductors. • <span class="hlt">Na</span>{sub <span class="hlt">2</span>}Ga{sub <span class="hlt">2</span>}GeS{sub 6</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li class="active"><span>18</span></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_18 --> <div id="page_19" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li class="active"><span>19</span></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="361"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29244481','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29244481"><span>Nanostructured <span class="hlt">Na</span><span class="hlt">2</span>Ti9O19 for Hybrid Sodium-Ion Capacitors with Excellent Rate Capability.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Bhat, Swetha S M; Babu, Binson; Feygenson, Mikhail; Neuefeind, Joerg C; Shaijumon, M M</p> <p>2018-01-10</p> <p>Herein, we report a new <span class="hlt">Na</span>-insertion electrode material, <span class="hlt">Na</span> <span class="hlt">2</span> Ti 9 O 19 , as a potential candidate for <span class="hlt">Na</span>-ion hybrid capacitors. We study the structural properties of nanostructured <span class="hlt">Na</span> <span class="hlt">2</span> Ti 9 O 19 , synthesized by a hydrothermal technique, upon electrochemical cycling vs <span class="hlt">Na</span>. Average and local structures of <span class="hlt">Na</span> <span class="hlt">2</span> Ti 9 O 19 are elucidated from neutron Rietveld refinement and pair distribution function (PDF), respectively, to investigate the initial discharge and charge events. Rietveld refinement reveals electrochemical cycling of <span class="hlt">Na</span> <span class="hlt">2</span> Ti 9 O 19 is driven by single-phase solid solution reaction during (de)sodiation without any major structural deterioration, keeping the average structure intact. Unit cell volume and lattice evolution on discharge process is inherently related to TiO 6 distortion and <span class="hlt">Na</span> ion perturbations, while the PDF reveals the deviation in the local structure after sodiation. Raman spectroscopy and X-ray photoelectron spectroscopy studies further corroborate the average and local structural behavior derived from neutron diffraction measurements. Also, <span class="hlt">Na</span> <span class="hlt">2</span> Ti 9 O 19 shows excellent <span class="hlt">Na</span>-ion kinetics with a capacitve nature of 86% at 1.0 mV s -1 , indicating that the material is a good anode candidate for a sodium-ion hybrid capacitor. A full cell hybrid <span class="hlt">Na</span>-ion capacitor is fabricated by using <span class="hlt">Na</span> <span class="hlt">2</span> Ti 9 O 19 as anode and activated porous carbon as cathode, which exhibits excellent electrochemical properties, with a maximum energy density of 54 Wh kg -1 and a maximum power density of 5 kW kg -1 . Both structural insights and electrochemical investigation suggest that <span class="hlt">Na</span> <span class="hlt">2</span> Ti 9 O 19 is a promising negative electrode for sodium-ion batteries and hybrid capacitors.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22476115-growth-mechanism-magnetic-electrochemical-properties-na-sub-mno-sub-nanorods-cathode-material-na-ion-batteries','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22476115-growth-mechanism-magnetic-electrochemical-properties-na-sub-mno-sub-nanorods-cathode-material-na-ion-batteries"><span>Growth mechanism and magnetic and electrochemical properties of <span class="hlt">Na</span>{sub 0.44}MnO{sub <span class="hlt">2</span>} nanorods as cathode material for <span class="hlt">Na</span>-ion batteries</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>Demirel, S.; Oz, E.; Altin, E.</p> <p></p> <p>Nanorods of <span class="hlt">Na</span>{sub 0.44}MnO{sub <span class="hlt">2</span>} are a promising cathode material for <span class="hlt">Na</span>-ion batteries due to their large surface area and single crystalline structure. We report the growth mechanism of <span class="hlt">Na</span>{sub 0.44}MnO{sub <span class="hlt">2</span>} nanorods via solid state synthesis and their physical properties. The structure and the morphology of the <span class="hlt">Na</span>{sub 0.44}MnO{sub <span class="hlt">2</span>} nanorods are investigated by X-ray diffraction (XRD), scanning and tunneling electron microscopy (SEM and TEM), and energy-dispersive X-ray (EDX) techniques. The growth mechanism of the rods is investigated and the effects of vapor pressure and partial melting of <span class="hlt">Na</span>-rich regions are discussed. The magnetic measurements show an antiferromagnetic phasemore » transition at 25 K and the μ{sub eff} is determined as 3.41 and 3.24 μ{sub B} from the χ–T curve and theoretical calculation, respectively. The electronic configuration and spin state of Mn{sup 3+} and Mn{sup 4+} are discussed in detail. The electrochemical properties of the cell fabricated using the nanorods are investigated and the peaks in the voltammogram are attributed to the diffusion of <span class="hlt">Na</span> ions from different sites. <span class="hlt">Na</span> intercalation process is explained by one and two Margules and van Laar models. - Highlights: • We synthesized <span class="hlt">Na</span>{sub 0.44}MnO{sub <span class="hlt">2</span>} nanorods via a simple solid state reaction technique. • Our studies show that excess <span class="hlt">Na</span> plays a crucial role in the nanorod formation. • Magnetization measurements show that Mn{sup 3+} ions are in LS and HS states. • The electrochemical properties of the cell fabricated using the nanorods are investigated. • <span class="hlt">Na</span> intercalation process is explained by one and two Margules and van Laar models.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018PhyB..535..157T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PhyB..535..157T"><span>Luminescence properties of <span class="hlt">Na</span><span class="hlt">2</span>Sr<span class="hlt">2</span>Al<span class="hlt">2</span>PO4Cl9:Sm3+ phosphor</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Tamboli, Sumedha; Shahare, D. I.; Dhoble, S. J.</p> <p>2018-04-01</p> <p>A series of Sm3+ ions doped <span class="hlt">Na</span><span class="hlt">2</span>Sr<span class="hlt">2</span>Al<span class="hlt">2</span>PO4Cl9 phosphors were synthesized via solid state synthesis method. Photoluminescence (PL) emission spectra were obtained by keeping excitation wavelength at 406 nm. Emission spectra show three emission peaks at 563 nm, 595 nm and 644 nm. The CIE chromaticity diagram shows emission colour of the phosphor in the orange-red region of the visible spectrum, indicating that the phosphor may be useful in preparing orange light-emitting diodes. <span class="hlt">Na</span><span class="hlt">2</span>Sr<span class="hlt">2</span>Al<span class="hlt">2</span>PO4Cl9:Sm3+ phosphors were irradiated by γ-rays from a 60Co source and β-rays from a 90Sr source. Their thermoluminescence (TL) glow curves were obtained by Nucleonix 1009I TL reader. TL Trapping parameters such as activation energy of trapped electrons and order of kinetics were obtained by using Chen's peak shape method, Glow curve fitting method and initial rise method.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2006AIPC..833...61O','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2006AIPC..833...61O"><span>Effects of <span class="hlt">Na</span>OH Concentration on CO<span class="hlt">2</span> Reduction via Hydrothermal Water</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Onoki, Takamasa; Takahashi, Hiro; Kori, Toshinari; Yamasaki, Nakamichi; Hashida, Toshiyuki</p> <p>2006-05-01</p> <p>The reductions of CO<span class="hlt">2</span> under hydrothermal conditions were investigated by using the micro autoclave (45cm3) lined with Hastelloy-C alloy. Sodium hydrogen carbonate (<span class="hlt">Na</span>HCO3) was used as a starting material. H<span class="hlt">2</span> gas was used as reducing agents. <span class="hlt">Na</span>HCO3 powder, H<span class="hlt">2</span> gas and water put into the autoclave simultaneously. The autoclave was heated upto 300°C by induction heater. In this study, effects of pH value of the <span class="hlt">Na</span>OH solution in the autoclave are investigated. Reaction products were analyzed with gas chromatographs (GC), liquid chromatographs (LC), X-ray diffractometor (XRD) and Scanning electron microscopy (SEM). The following things were showed in this research: CO<span class="hlt">2</span> was reducted to HCOO- and CH4 at high conversion ratio under hydrothermal conditions. HCOO- was formed at high selectivity using Hastelloy-C reactor in the alkaline solution with Raney Ni catalyst. Raney Ni was exellent methanation catalyst, and CH4 formation progressed via HCO3-, not via CO. It is cleared that the <span class="hlt">Na</span>OH solution in the autoclave should be kept pH value 11.0 for the highest conversion ratio from CO<span class="hlt">2</span> to useful carbonic compounds (CH4, HCOO-).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017APS..MARF38003Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017APS..MARF38003Z"><span>Polaronic and ionic conduction in <span class="hlt">Na</span>MnO<span class="hlt">2</span>: influence of native point defects</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zhu, Zhen; Peelaers, Hartwin; van de Walle, Chris G.</p> <p></p> <p>Layered <span class="hlt">Na</span>MnO<span class="hlt">2</span> has promising applications as a cathode material for sodium ion batteries. We will discuss strategies to improve the electrical performance of <span class="hlt">Na</span>MnO<span class="hlt">2</span>, including how to optimize the conditions of synthesis and how impurity doping affects the performance. Using hybrid density functional theory, we explored the structural, electronic, and defect properties of bulk <span class="hlt">Na</span>MnO<span class="hlt">2</span>. It is antiferromagnetic in the ground state with a band gap of 3.75 eV. Small hole and electron polarons can form in the bulk either through self-trapping or adjacent to point defects. We find that both <span class="hlt">Na</span> and Mn vacancies are shallow acceptors with the induced holes trapped as small polarons, while O vacancies are deep defect centers. Cation antisites, especially Mn<span class="hlt">Na</span>, are found to have low formation energies. As a result, we expect that Mn<span class="hlt">Na</span> exists in as-grown <span class="hlt">Na</span>MnO<span class="hlt">2</span> in moderate concentrations, rather than forming only at a later stage of the charging process, at which point it causes undesirable structural phase transitions. Both electronic conduction, via polaron hopping, and ionic conduction, through VNa migration, are significantly affected by the presence of point defects. This work was supported by DOE.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1193224-ordered-disordered-polymorphs-na-ni2-honeycomb-ordered-cathodes-na-ion-batteries','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1193224-ordered-disordered-polymorphs-na-ni2-honeycomb-ordered-cathodes-na-ion-batteries"><span>Ordered and disordered polymorphs of <span class="hlt">Na</span>(Ni <span class="hlt">2</span>/3Sb 1/3)O₂: Honeycomb-ordered cathodes for <span class="hlt">Na</span>-ion batteries</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Ma, Jeffrey; Wu, Lijun; Bo, Shou -Hang; ...</p> <p>2015-04-14</p> <p><span class="hlt">Na</span>-ion batteries are appealing alternatives to Li-ion battery systems for large-scale energy storage applications in which elemental cost and abundance are important. Although it is difficult to find <span class="hlt">Na</span>-ion batteries which achieve substantial specific capacities at voltages above 3 V (vs Na⁺/<span class="hlt">Na</span>), the honeycomb-layered compound <span class="hlt">Na</span>(Ni <span class="hlt">2</span>/3Sb 1/3)O₂ can deliver up to 130 mAh/g of capacity at voltages above 3 V with this capacity concentrated in plateaus at 3.27 and 3.64 V. Comprehensive crystallographic studies have been carried out in order to understand the role of disorder in this system which can be prepared in both “disordered” and “ordered” forms,more » depending on the synthesis conditions. The average structure of <span class="hlt">Na</span>(Ni <span class="hlt">2</span>/3Sb 1/3)O₂ is always found to adopt an O3-type stacking sequence, though different structures for the disordered (R3¯ m, #166, a = b = 3.06253(3) Å and c = 16.05192(7) Å) and ordered variants ( C<span class="hlt">2</span>/m, #12, a = 5.30458(1) Å, b = 9.18432(1) Å, c = 5.62742(1) Å and β = 108.2797(<span class="hlt">2</span>)°) are demonstrated through the combined Rietveld refinement of synchrotron X-ray and time-of-flight neutron powder diffraction data. However, pair distribution function studies find that the local structure of disordered <span class="hlt">Na</span>(Ni <span class="hlt">2</span>/3Sb 1/3)O₂ is more correctly described using the honeycomb-ordered structural model, and solid state NMR studies confirm that the well-developed honeycomb ordering of Ni and Sb cations within the transition metal layers is indistinguishable from that of the ordered phase. The disorder is instead found to mainly occur perpendicular to the honeycomb layers with an observed coherence length of not much more than 1 nm seen in electron diffraction studies. When the <span class="hlt">Na</span> environment is probed through ²³<span class="hlt">Na</span> solid state NMR, no evidence is found for prismatic <span class="hlt">Na</span> environments, and a bulk diffraction analysis finds no evidence of conventional stacking faults. The lack of long range coherence is instead attributed to disorder among</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/20447473','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/20447473"><span>A new sol-gel process for producing <span class="hlt">Na</span>(<span class="hlt">2</span>)O-containing bioactive glass ceramics.</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-Zhi; Li, Yuan; Jin, Li-Yu; Quinn, Julian M W; Komesaroff, Paul A</p> <p>2010-10-01</p> <p>The sol-gel process of producing SiO(<span class="hlt">2</span>)-CaO bioactive glasses is well established, but problems remain with the poor mechanical properties of the amorphous form and the bioinertness of its crystalline counterpart. These properties may be improved by incorporating <span class="hlt">Na</span>(<span class="hlt">2</span>)O into bioactive glasses, which can result in the formation of a hard yet biodegradable crystalline phase from bioactive glasses when sintered. However, production of <span class="hlt">Na</span>(<span class="hlt">2</span>)O-containing bioactive glasses by sol-gel methods has proved to be difficult. This work reports a new sol-gel process for the production of <span class="hlt">Na</span>(<span class="hlt">2</span>)O-containing bioactive glass ceramics, potentially enabling their use as medical implantation materials. Fine powders of 45S5 (a <span class="hlt">Na</span>(<span class="hlt">2</span>)O-containing composition) glass ceramic have for the first time been successfully synthesized using the sol-gel technique in aqueous solution under ambient conditions, with the mean particle size being approximately 5 microm. A comparative study of sol-gel derived S70C30 (a <span class="hlt">Na</span>(<span class="hlt">2</span>)O-free composition) and 45S5 glass ceramic materials revealed that the latter possesses a number of features desirable in biomaterials used for bone tissue engineering, including (i) the crystalline phase <span class="hlt">Na</span>(<span class="hlt">2</span>)Ca(<span class="hlt">2</span>)Si(3)O(9) that couples good mechanical strength with satisfactory biodegradability, (ii) formation of hydroxyapatite, which may promote good bone bonding and (iii) cytocompatibility. In contrast, the sol-gel derived S70C30 glass ceramic consisted of a virtually inert crystalline phase CaSiO(3). Moreover, amorphous S70C30 largely transited to CaCO(3) with minor hydroxyapatite when immersed in simulated body fluid under standard tissue culture conditions. In conclusion, sol-gel derived <span class="hlt">Na</span>(<span class="hlt">2</span>)O-containing glass ceramics have significant advantages over related <span class="hlt">Na</span>(<span class="hlt">2</span>)O-free materials, having a greatly improved combination of mechanical capability and biological absorbability. 2010 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/1048960-structure-temperature-dependent-phase-transitions-lead-free-bi1-bi1-k0-piezoceramics','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/1048960-structure-temperature-dependent-phase-transitions-lead-free-bi1-bi1-k0-piezoceramics"><span>Structure and temperature-dependent phase transitions of lead-free Bi 1/<span class="hlt">2</span><span class="hlt">Na</span> 1/<span class="hlt">2</span>TiO 3-Bi 1/<span class="hlt">2</span>K 1/<span class="hlt">2</span>TiO 3-K 0.5<span class="hlt">Na</span> 0.5NbO 3 piezoceramics</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>Anton, Eva-Maria; Schmitt, Ljubomira Ana; Hinterstein, Manuel</p> <p>2014-05-28</p> <p>Structure and phase transitions of (1-y)((1-x)Bi 1/<span class="hlt">2</span><span class="hlt">Na</span> 1/<span class="hlt">2</span>TiO 3-xBi 1/<span class="hlt">2</span>K 1/<span class="hlt">2</span>TiO 3)-yK 0.5<span class="hlt">Na</span> 0.5NbO 3 (x; y) piezoceramics (0.1 ≤ x ≤ 0.4; 0 ≤ y ≤ 0.05) were investigated by transmission electron microscopy, neutron diffraction, temperature-dependent x-ray diffraction, and Raman spectroscopy. The local crystallographic structure at room temperature (RT) does not change by adding K 0.5<span class="hlt">Na</span> 0.5NbO 3 to Bi 1/<span class="hlt">2</span><span class="hlt">Na</span> 1/<span class="hlt">2</span>TiO 3-xBi 1/<span class="hlt">2</span>K 1/<span class="hlt">2</span>TiO 3 for x = 0.<span class="hlt">2</span> and 0.4. The average crystal structure and microstructure on the other hand develop from mainly long-range polar order with ferroelectric domains to short-range order with polar nanoregions displaying amore » more pronounced relaxor character. The (0.1; 0) and (0.1; 0.02) compositions exhibit monoclinic Cc space group symmetry, which transform into Cc + P4bm at 185 and 130 °C, respectively. This high temperature phase is stable at RT for the morphotropic phase boundary compositions of (0.1; 0.05) and all compositions with x = 0.<span class="hlt">2</span>. For the compositions of (0.1; 0) and (0.1; 0.02), local structural changes on heating are evidenced by Raman; for all other compositions, changes in the long-range average crystal structure were observed.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/5740766-sodium-relations-desert-plants-differential-effects-nacl-na-sub-so-sub-growth-composition-atriplex-hymenelytra-desert-holly','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/5740766-sodium-relations-desert-plants-differential-effects-nacl-na-sub-so-sub-growth-composition-atriplex-hymenelytra-desert-holly"><span>Sodium relations in desert plants: 8. Differential effects of <span class="hlt">Na</span>Cl and <span class="hlt">Na</span>/sub <span class="hlt">2</span>/SO/sub 4/ on growth and composition of Atriplex hymenelytra (desert holly)</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>Soufi, S.M.; Wallace, A.</p> <p>1982-07-01</p> <p>Maximum growth over a period of 3 months of Atriplex hymenelytra (Torr.) Wats. (desert holly) in solution culture was obtained when the nutrient solution contained 5 x 10/sup -<span class="hlt">2</span>/ N <span class="hlt">Na</span>Cl. Sodium concentratons in leaves at maximum yield was 7.88% and that of Cl was also 7.88%. In the presence of 10/sup -<span class="hlt">2</span>/ N <span class="hlt">Na</span>/sub <span class="hlt">2</span>/SO/sub 4/, there was much less growth than with 10/sup -<span class="hlt">2</span>/ N <span class="hlt">Na</span>Cl. The highest <span class="hlt">Na</span>Cl level depressed levels of K, Ca, and Mg in leaves, stems, and roots. The highest <span class="hlt">Na</span>Cl level also decreased levels of micronutrients in many of the plants.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/966546','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/966546"><span>The Origin of Refractory Minerals in Comet 81P/Wild <span class="hlt">2</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>Chi, M; Ishii, H A; Simon, S B</p> <p>2008-11-20</p> <p>Refractory Ti-bearing minerals in the calcium-, aluminium-rich inclusion (<span class="hlt">CAI</span>) Inti, recovered from the comet 81P/Wild <span class="hlt">2</span> sample, were examined using analytical (scanning) transmission electron microscopy (STEM) methods including imaging, nanodiffraction, energy dispersive spectroscopy (EDX) and electron energy loss spectroscopy (EELS). Inti fassaite (Ca(Mg,Ti,Al)(Si,Al){sub <span class="hlt">2</span>}O{sub 6}) was found to have a Ti{sup 3+}/Ti{sup 4+} ratio of <span class="hlt">2</span>.0 {+-} 0.<span class="hlt">2</span>, consistent with fassaite in other solar system <span class="hlt">CAIs</span>. The oxygen fugacity (log f{sub O{sub <span class="hlt">2</span>}}) of formation estimated from this ratio, assuming equilibration among phases at 1509K, is -19.4 {+-} 1.3. This value is near the canonical solar nebula value (-18.1 {+-}more » 0.3) and in close agreement with that reported for fassaite-bearing Allende <span class="hlt">CAIs</span> (-19.8 {+-} 0.9) by other researchers using the same assumptions. Nanocrystals of osbornite (Ti(V)N), <span class="hlt">2</span>-40 nm in diameter, are embedded as inclusions within anorthite, spinel and diopside in Inti. Vanadium is heterogeneously distributed within some osbornite crystals. Compositions range from pure TiN to Ti{sub 0.36}V{sub 0.64}N. The possible presence of oxide and carbide in solid solution with the osbornite was evaluated. The osbornite may contain O but does not contain C. The presence of osbornite, likely a refractory early condensate, together with the other refractory minerals in Inti, indicates that the parent comet contains solids that condensed closer to the proto-sun than the distance at which the parent comet itself accreted. The estimated oxygen fugacity and the reported isotopic and chemical compositions are consistent with Inti originating in the inner solar system as opposed to it being a surviving <span class="hlt">CAI</span> from an extrasolar source. These results provide insight for evaluating the validity of models of radial mass transport dynamics in the early solar system. The oxidation environments inferred for the Inti mineral assemblage are inconsistent with an X</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/12844514','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/12844514"><span>Mechanisms for monovalent cation-dependent depletion of intracellular Mg<span class="hlt">2</span>+:<span class="hlt">Na</span>(+)-independent Mg<span class="hlt">2</span>+ pathways in guinea-pig smooth muscle.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Nakayama, Shinsuke; Nomura, Hideki; Smith, Lorraine M; Clark, Joseph F; Uetani, Tadayuki; Matsubara, Tatsuaki</p> <p>2003-09-15</p> <p>It has been suggested that magnesium deficiency is correlated with many diseases. 31P NMR experiments were carried out in order to investigate the effects of <span class="hlt">Na</span>+ substitution on Mg<span class="hlt">2</span>+ depletion in smooth muscle under divalent cation-free conditions. In the taenia of guinea-pig caeci, the intracellular free Mg<span class="hlt">2</span>+ concentration ([Mg<span class="hlt">2</span>+]i) was estimated from the chemical shifts of (1) the beta-ATP peak alone and (<span class="hlt">2</span>) beta- and gamma-ATP peaks. Both estimations indicated that [Mg<span class="hlt">2</span>+]i decreased only very slowly in Mg(<span class="hlt">2</span>+)-free, Ca(<span class="hlt">2</span>+)-free solutions in which <span class="hlt">Na</span>+ was substituted with large cations such as NMDG (N-methyl-D-glucamine) and choline. Furthermore, the measurements of tension development supported the suggestion of preservation of intracellular Mg<span class="hlt">2</span>+ with NMDG substitution. Substituting extracellular <span class="hlt">Na</span>+ with the small cation, Li+, also shifted the beta-ATP peak towards a lower frequency, but the frequency shift was significantly less than that seen upon <span class="hlt">Na</span>+ substitution with K+. The estimated [Mg<span class="hlt">2</span>+]i depletion was, however, comparable with that seen after <span class="hlt">Na</span>+ substitution with K+ using the titration curves of metal-free and Mg(<span class="hlt">2</span>+)-bound ATP obtained in Li(+)-based model solutions. It was concluded that Mg<span class="hlt">2</span>+ rapidly decreases only when small cations were the major electrolyte of the extracellular medium. <span class="hlt">Na</span>+ substitutions with NMDG, choline or Li+ had little effect on intracellular ATP concentration after 100 min treatment.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29775274','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29775274"><span>Oxide-Based Composite Electrolytes Using <span class="hlt">Na</span>3Zr<span class="hlt">2</span>Si<span class="hlt">2</span>PO12/<span class="hlt">Na</span>3PS4 Interfacial Ion Transfer.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Noi, Kousuke; Nagata, Yuka; Hakari, Takashi; Suzuki, Kenji; Yubuchi, So; Ito, Yusuke; Sakuda, Atsushi; Hayashi, Akitoshi; Tatsumisago, Masahiro</p> <p>2018-05-31</p> <p>All-solid-state sodium batteries using <span class="hlt">Na</span> 3 Zr <span class="hlt">2</span> Si <span class="hlt">2</span> PO 12 (NASICON) solid electrolytes are promising candidates for safe and low-cost advanced rechargeable battery systems. Although NASICON electrolytes have intrinsically high sodium-ion conductivities, their high sintering temperatures interfere with the immediate development of high-performance batteries. In this work, sintering-free NASICON-based composites with <span class="hlt">Na</span> 3 PS 4 (NPS) glass ceramics were prepared to combine the high grain-bulk conductivity of NASICON and the interfacial formation ability of NPS. Before the composite preparation, the NASICON/NPS interfacial resistance was investigated by modeling the interface between the NASICON sintered ceramic and the NPS glass thin film. The interfacial ion-transfer resistance was very small above room temperature; the area-specific resistances at 25 and 100 °C were 15.8 and 0.40 Ω cm <span class="hlt">2</span> , respectively. On the basis of this smooth ion transfer, NASICON-rich (70-90 wt %) NASICON-NPS composite powders were prepared by ball-milling fine powders of each component. The composite powders were well-densified by pressing at room temperature. Scanning electron microscopy observation showed highly dispersed sub-micrometer NASICON grains in a dense NPS matrix to form closed interfaces between the oxide and sulfide solid electrolytes. The composite green (unfired) compacts with 70 and 80 wt % NASICON exhibited high total conductivities at 100 °C of 1.1 × 10 -3 and 6.8 × 10 -4 S cm -1 , respectively. An all-solid-state <span class="hlt">Na</span> 15 Sn 4 /TiS <span class="hlt">2</span> cell was constructed using the 70 wt % NASICON composite electrolyte by the uniaxial pressing of the powder materials, and its discharge properties were evaluated at 100 °C. The cell showed the reversible capacities of about 120 mAh g -1 under the current density of 640 μA cm -<span class="hlt">2</span> . The prepared oxide-based composite electrolytes were thus successfully applied in all-solid-state sodium rechargeable batteries without sintering.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012RJPCA..86.2076R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012RJPCA..86.2076R"><span>Interactions in L-phenylalanine/L-leucine/L-glutamic Acid/L-proline + <span class="hlt">2</span> M aqueous <span class="hlt">Na</span>Cl/<span class="hlt">2</span> M <span class="hlt">Na</span>NO3 systems at different temperatures</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Riyazuddeen, Imran Khan; Afrin, Sadaf</p> <p>2012-12-01</p> <p>Density (ρ) and speed of sound ( u) in <span class="hlt">2</span> M aqueous <span class="hlt">Na</span>Cl and <span class="hlt">2</span> M <span class="hlt">Na</span>NO3 solutions of amino acids: L-phenylalanine, L-leucine, L-glutamic acid, and L-proline have been measured for several molal concentrations of amino acids at different temperatures. The ρ and u data have been used to calculate the values of isothermal compressibility and internal pressure at different temperatures. The trends of variations of κ T and P i with an increase in molal concentration of amino acid and temperature have been discussed in terms of solute-solvent and solute-solute interactions in the systems.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3171941','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3171941"><span>Nedd4-<span class="hlt">2</span> Modulates Renal <span class="hlt">Na</span>+-Cl− Cotransporter via the Aldosterone-SGK1-Nedd4-<span class="hlt">2</span> Pathway</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Arroyo, Juan Pablo; Lagnaz, Dagmara; Ronzaud, Caroline; Vázquez, Norma; Ko, Benjamin S.; Moddes, Lauren; Ruffieux-Daidié, Dorothée; Hausel, Pierrette; Koesters, Robert; Yang, Baoli; Stokes, John B.; Hoover, Robert S.</p> <p>2011-01-01</p> <p>Regulation of renal <span class="hlt">Na</span>+ transport is essential for controlling blood pressure, as well as <span class="hlt">Na</span>+ and K+ homeostasis. Aldosterone stimulates <span class="hlt">Na</span>+ reabsorption by the <span class="hlt">Na</span>+-Cl− cotransporter (NCC) in the distal convoluted tubule (DCT) and by the epithelial <span class="hlt">Na</span>+ channel (ENaC) in the late DCT, connecting tubule, and collecting duct. Aldosterone increases ENaC expression by inhibiting the channel's ubiquitylation and degradation; aldosterone promotes serum-glucocorticoid-regulated kinase SGK1-mediated phosphorylation of the ubiquitin-protein ligase Nedd4-<span class="hlt">2</span> on serine 328, which prevents the Nedd4-<span class="hlt">2</span>/ENaC interaction. It is important to note that aldosterone increases NCC protein expression by an unknown post-translational mechanism. Here, we present evidence that Nedd4-<span class="hlt">2</span> coimmunoprecipitated with NCC and stimulated NCC ubiquitylation at the surface of transfected HEK293 cells. In Xenopus laevis oocytes, coexpression of NCC with wild-type Nedd4-<span class="hlt">2</span>, but not its catalytically inactive mutant, strongly decreased NCC activity and surface expression. SGK1 prevented this inhibition in a kinase-dependent manner. Furthermore, deficiency of Nedd4-<span class="hlt">2</span> in the renal tubules of mice and in cultured mDCT15 cells upregulated NCC. In contrast to ENaC, Nedd4-<span class="hlt">2</span>-mediated inhibition of NCC did not require the PY-like motif of NCC. Moreover, the mutation of Nedd4-<span class="hlt">2</span> at either serine 328 or 222 did not affect SGK1 action, and mutation at both sites enhanced Nedd4-<span class="hlt">2</span> activity and abolished SGK1-dependent inhibition. Taken together, these results suggest that aldosterone modulates NCC protein expression via a pathway involving SGK1 and Nedd4-<span class="hlt">2</span> and provides an explanation for the well-known aldosterone-induced increase in NCC protein expression. PMID:21852580</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://ajpregu.physiology.org/content/280/6/R1844','USGSPUBS'); return false;" href="http://ajpregu.physiology.org/content/280/6/R1844"><span>Gill <span class="hlt">Na</span>+-K+-<span class="hlt">2</span>Cl- cotransporter abundance and location in Atlantic salmon: Effects of seawater and smolting</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Pelis, Ryan M.; Zydlewski, Joseph D.; McCormick, Stephen D.</p> <p>2001-01-01</p> <p><span class="hlt">Na</span>+-K+-<span class="hlt">2</span>Cl−cotransporter abundance and location was examined in the gills of Atlantic salmon (Salmo salar) during seawater acclimation and smolting. Western blots revealed three bands centered at 285, 160, and 120 kDa. The <span class="hlt">Na</span>+-K+-<span class="hlt">2</span>Cl−cotransporter was colocalized with <span class="hlt">Na</span>+-K+-ATPase to chloride cells on both the primary filament and secondary lamellae. Parr acclimated to 30 parts per thousand seawater had increased gill <span class="hlt">Na</span>+-K+-<span class="hlt">2</span>Cl− cotransporter abundance, large and numerous <span class="hlt">Na</span>+-K+-<span class="hlt">2</span>Cl− cotransporter immunoreactive chloride cells on the primary filament, and reduced numbers on the secondary lamellae. Gill <span class="hlt">Na</span>+-K+-<span class="hlt">2</span>Cl− cotransporter levels were low in presmolts (February) and increased 3.3-fold in smolts (May), coincident with elevated seawater tolerance. Cotransporter levels decreased below presmolt values in postsmolts in freshwater (June). The size and number of immunoreactive chloride cells on the primary filament increased threefold during smolting and decreased in postsmolts. Gill <span class="hlt">Na</span>+-K+-ATPase activity and <span class="hlt">Na</span>+-K+-<span class="hlt">2</span>Cl− cotransporter abundance increased in parallel during both seawater acclimation and smolting. These data indicate a direct role of the <span class="hlt">Na</span>+-K+-<span class="hlt">2</span>Cl− cotransporter in salt secretion by gill chloride cells of teleost fish.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/20843809','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/20843809"><span>Histidine residues in the <span class="hlt">Na</span>+-coupled ascorbic acid transporter-<span class="hlt">2</span> (SVCT<span class="hlt">2</span>) are central regulators of SVCT<span class="hlt">2</span> function, modulating pH sensitivity, transporter kinetics, <span class="hlt">Na</span>+ cooperativity, conformational stability, and subcellular localization.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Ormazabal, Valeska; Zuñiga, Felipe A; Escobar, Elizabeth; Aylwin, Carlos; Salas-Burgos, Alexis; Godoy, Alejandro; Reyes, Alejandro M; Vera, Juan Carlos; Rivas, Coralia I</p> <p>2010-11-19</p> <p><span class="hlt">Na</span>(+)-coupled ascorbic acid transporter-<span class="hlt">2</span> (SVCT<span class="hlt">2</span>) activity is impaired at acid pH, but little is known about the molecular determinants that define the transporter pH sensitivity. SVCT<span class="hlt">2</span> contains six histidine residues in its primary sequence, three of which are exofacial in the transporter secondary structure model. We used site-directed mutagenesis and treatment with diethylpyrocarbonate to identify histidine residues responsible for SVCT<span class="hlt">2</span> pH sensitivity. We conclude that five histidine residues, His(109), His(203), His(206), His(269), and His(413), are central regulators of SVCT<span class="hlt">2</span> function, participating to different degrees in modulating pH sensitivity, transporter kinetics, <span class="hlt">Na</span>(+) cooperativity, conformational stability, and subcellular localization. Our results are compatible with a model in which (i) a single exofacial histidine residue, His(413), localized in the exofacial loop IV that connects transmembrane helices VII-VIII defines the pH sensitivity of SVCT<span class="hlt">2</span> through a mechanism involving a marked attenuation of the activation by <span class="hlt">Na</span>(+) and loss of <span class="hlt">Na</span>(+) cooperativity, which leads to a decreased V(max) without altering the transport K(m); (ii) exofacial histidine residues His(203), His(206), and His(413) may be involved in maintaining a functional interaction between exofacial loops II and IV and influence the general folding of the transporter; (iii) histidines 203, 206, 269, and 413 affect the transporter kinetics by modulating the apparent transport K(m); and (iv) histidine 109, localized at the center of transmembrane helix I, might be fundamental for the interaction of SVCT<span class="hlt">2</span> with the transported substrate ascorbic acid. Thus, histidine residues are central regulators of SVCT<span class="hlt">2</span> function.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016MMTB...47.2447Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016MMTB...47.2447Y"><span>Crystallization Behavior and Heat Transfer of Fluorine-Free Mold Fluxes with Different <span class="hlt">Na</span><span class="hlt">2</span>O Concentration</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yang, Jian; Zhang, Jianqiang; Sasaki, Yasushi; Ostrovski, Oleg; Zhang, Chen; Cai, Dexiang; Kashiwaya, Yoshiaki</p> <p>2016-08-01</p> <p>In this study, the crystallization behavior and heat transfer of CaO-SiO<span class="hlt">2</span>-<span class="hlt">Na</span><span class="hlt">2</span>O-B<span class="hlt">2</span>O3-TiO<span class="hlt">2</span>-Al<span class="hlt">2</span>O3-MgO-Li<span class="hlt">2</span>O fluorine-free mold fluxes with different <span class="hlt">Na</span><span class="hlt">2</span>O contents (5 to 11 mass pct) were studied using single/double hot thermocouple technique (SHTT/DHTT) and infrared emitter technique (IET), respectively. Continuous cooling transformation (CCT) and time-temperature transformation (TTT) diagrams constructed using SHTT showed that crystallization temperature increased and incubation time shortened with the increase of <span class="hlt">Na</span><span class="hlt">2</span>O concentration, indicating an enhanced crystallization tendency. The crystallization process of mold fluxes in the temperature field simulating the casting condition was also investigated using DHTT. X-ray diffraction (XRD) analysis of the quenched mold fluxes showed that the dominant phase changed from CaSiO3 to Ca11Si4B<span class="hlt">2</span>O22 with the increasing concentration of <span class="hlt">Na</span><span class="hlt">2</span>O. The heat transfer examined by IET showed that the increase of <span class="hlt">Na</span><span class="hlt">2</span>O concentration reduced the responding heat flux when <span class="hlt">Na</span><span class="hlt">2</span>O was lower than 9 mass pct but the further increase of <span class="hlt">Na</span><span class="hlt">2</span>O to 11 mass pct enhanced the heat flux. The correlation between crystallinity and heat transfer was discussed in terms of crystallization tendency and crystal morphology.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017MRE.....4j5204Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017MRE.....4j5204Z"><span>Effect of increasing concentration of <span class="hlt">Na</span><span class="hlt">2</span>O on structural, elastic and optical properties of (90  -  x)GeO<span class="hlt">2</span>-x<span class="hlt">Na</span><span class="hlt">2</span>O-10PbO glass system in the germanate anomaly region</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zainudin, C. N.; Hisam, R.; Yusof, M. I. M.; Yahya, A. K.; Halimah, M. K.</p> <p>2017-10-01</p> <p>Ternary germanate glasses (90  -  x)GeO<span class="hlt">2</span>-x<span class="hlt">Na</span><span class="hlt">2</span>O-10PbO (x  =  10-30 mol%) have been prepared by the melt-quenching method. Density, ρ increased with <span class="hlt">Na</span><span class="hlt">2</span>O content up to maxima at 20 mol% while molar volume, V a showed an opposite trend to the density, with a minima at 20 mol% of <span class="hlt">Na</span><span class="hlt">2</span>O content indicating the presence of the germanate anomaly. Ultrasonic velocity measurements showed both longitudinal, v l and shear, v s velocities increased up to 20 mol% before decreasing with further addition of <span class="hlt">Na</span><span class="hlt">2</span>O. Independent longitudinal, L and shear, G moduli along with Young’s modulus, Y, mean sound velocity, v m, Debye temperature, θ D, and hardness, H recorded maximum values at 20 mol% of <span class="hlt">Na</span><span class="hlt">2</span>O content which were suggested to be related to the germanate anomaly. Structural modification occurring due to conversion of six-membered GeO4 rings to three-membered rings of GeO4 changed bond density and compactness of the glass systems and caused the increase in rigidity and stiffness of the glasses. Beyond 20 mol% of <span class="hlt">Na</span><span class="hlt">2</span>O, the decrease in the elastic moduli was due to depolymerization of the glass network. Meanwhile, optical energy gap, E opt exhibited a minima at 20 mol% whereas Urbach energy, E U and refractive index, n showed a maxima at the same concentration, thereby indicating variation in polarizability due to changes in concentration of bridging and non-bridging oxygen.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26765283','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26765283"><span>A mixed iron-manganese based pyrophosphate cathode, <span class="hlt">Na</span><span class="hlt">2</span>Fe0.5Mn0.5P<span class="hlt">2</span>O7, for rechargeable sodium ion batteries.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Shakoor, Rana A; Park, Chan Sun; Raja, Arsalan A; Shin, Jaeho; Kahraman, Ramazan</p> <p>2016-02-07</p> <p>The development of secondary batteries based on abundant and cheap elements is vital. Among various alternatives to conventional lithium-ion batteries, sodium-ion batteries (SIBs) are promising due to the abundant resources and low cost of sodium. While there are many challenges associated with the SIB system, cathode is an important factor in determining the electrochemical performance of this battery system. Accordingly, ongoing research in the field of SIBs is inclined towards the development of safe, cost effective cathode materials having improved performance. In particular, pyrophosphate cathodes have recently demonstrated decent electrochemical performance and thermal stability. Herein, we report the synthesis, electrochemical properties, and thermal behavior of a novel <span class="hlt">Na</span><span class="hlt">2</span>Fe0.5Mn0.5P<span class="hlt">2</span>O7 cathode for SIBs. The material was synthesized through a solid state process. The structural analysis reveals that the mixed substitution of manganese and iron has resulted in a triclinic crystal structure (P1[combining macron] space group). Galvanostatic charge/discharge measurements indicate that <span class="hlt">Na</span><span class="hlt">2</span>Fe0.5Mn0.5P<span class="hlt">2</span>O7 is electrochemically active with a reversible capacity of ∼80 mA h g(-1) at a C/20 rate with an average redox potential of 3.<span class="hlt">2</span> V. (vs. <span class="hlt">Na/Na</span>(+)). It is noticed that 84% of initial capacity is preserved over 90 cycles showing promising cyclability. It is also noticed that the rate capability of <span class="hlt">Na</span><span class="hlt">2</span>Fe0.5Mn0.5P<span class="hlt">2</span>O7 is better than <span class="hlt">Na</span><span class="hlt">2</span>MnP<span class="hlt">2</span>O7. Ex situ and CV analyses indicate that <span class="hlt">Na</span><span class="hlt">2</span>Fe0.5Mn0.5P<span class="hlt">2</span>O7 undergoes a single phase reaction rather than a biphasic reaction due to different <span class="hlt">Na</span> coordination environment and different <span class="hlt">Na</span> site occupancy when compared to other pyrophosphate materials (<span class="hlt">Na</span><span class="hlt">2</span>FeP<span class="hlt">2</span>O7 and <span class="hlt">Na</span><span class="hlt">2</span>MnP<span class="hlt">2</span>O7). Thermogravimetric analysis (25-550 °C) confirms good thermal stability of <span class="hlt">Na</span><span class="hlt">2</span>Fe0.5Mn0.5P<span class="hlt">2</span>O7 with only <span class="hlt">2</span>% weight loss. Owing to promising electrochemical properties and decent thermal stability, <span class="hlt">Na</span><span class="hlt">2</span>Fe0.5Mn0.5P<span class="hlt">2</span>O7, can be an attractive cathode for SIBs.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014JPhCS.507a2037T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014JPhCS.507a2037T"><span>Iron isotope effect in the iron arsenide superconductor (Ca0.4<span class="hlt">Na</span>0.6)Fe<span class="hlt">2</span>As<span class="hlt">2</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Tsuge, Y.; Nishio, T.; Iyo, A.; Tanaka, Y.; Eisaki, H.</p> <p>2014-05-01</p> <p>We report a new sample synthesis technique for polycrystalline (Ca1-xNax)Fe<span class="hlt">2</span>As<span class="hlt">2</span> (0<x<=0.6) and measurements of an iron isotope effect in optimally doped (Ca0.4<span class="hlt">Na</span>0.6)Fe<span class="hlt">2</span>As<span class="hlt">2</span> with three types of iron isotopes (54Fe, natural Fe, and 57Fe). We synthesized isotope samples carefully not to give rise to a difference in the <span class="hlt">Na</span> content x between different isotope samples, which becomes potentially a factor for an extrinsic difference in the superconducting transition temperature Tc between those samples. No significant difference in lattice parameters between those samples is shown by measurements of powder x-ray diffraction (XRD), implying that the <span class="hlt">Na</span> content in samples is well-controlled. Our estimate of the iron isotope coefficient αFe defined by -d InTc/d lnMFe, where MFe is the iron isotope mass, is -0.19. These indicate that in (Ca0.4<span class="hlt">Na</span>0.6)Fe<span class="hlt">2</span>As<span class="hlt">2</span>, the iron isotope coefficient becomes definitely negative. We discuss the implications of this fact, considering previous measurements of an iron isotope effect in different iron-based superconductors.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li class="active"><span>19</span></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_19 --> <div id="page_20" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li class="active"><span>20</span></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="381"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29870231','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29870231"><span>β-<span class="hlt">Na</span><span class="hlt">2</span>TeO4: Phase Transition from an Orthorhombic to a Monoclinic Form. Reversible CO<span class="hlt">2</span> Capture.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Galven, Cyrille; Pagnier, Thierry; Rosman, Noël; Le Berre, Françoise; Crosnier-Lopez, Marie-Pierre</p> <p>2018-06-18</p> <p>The present work concerns the tellurate <span class="hlt">Na</span> <span class="hlt">2</span> TeO 4 which has a 1D structure and could then present a CO <span class="hlt">2</span> capture ability. It has been synthesized in a powder form via a solid-state reaction and structurally characterized by thermal X-ray diffraction experiments, Raman spectroscopy, and differential scanning calorimetry. The room temperature structure corresponds to the β-<span class="hlt">Na</span> <span class="hlt">2</span> TeO 4 orthorhombic form, and we show that it undergoes a reversible structural transition near 420 °C toward a monoclinic system. Ab initio computations were also performed on the room temperature structure, the Raman vibration modes calculated, and a normal mode attribution proposed. In agreement with our expectations, this sodium oxide is able to trap CO <span class="hlt">2</span> by a two-step mechanism: <span class="hlt">Na</span> + /H + exchange and carbonation of the released sodium as <span class="hlt">Na</span>HCO 3 . This capture is reversible since CO <span class="hlt">2</span> can be released upon heating by recombination of the mother phase.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1372403-floating-zone-growth-na0-single-crystals','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1372403-floating-zone-growth-na0-single-crystals"><span>Floating zone growth of α-<span class="hlt">Na</span> 0.90MnO <span class="hlt">2</span> single crystals</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Dally, Rebecca; Clement, Raphaele J.; Chisnell, Robin; ...</p> <p>2016-12-03</p> <p>Here, single crystal growth of α-<span class="hlt">Na</span> xMnO <span class="hlt">2</span> (x=0.90) is reported via the floating zone technique. The conditions required for stable growth and intergrowth-free crystals are described along with the results of trials under alternate growth atmospheres. Chemical and structural characterizations of the resulting α-<span class="hlt">Na</span> 0.90MnO <span class="hlt">2</span> crystals are performed using ICP-AES NMR, XANES, XPS, and neutron diffraction measurements. As a layered transition metal oxide with large ionic mobility and strong correlation effects, α-<span class="hlt">Na</span> xMnO <span class="hlt">2</span> is of interest to many communities, and the implications of large volume, high purity, single crystal growth are discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1994JSSCh.110...74M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1994JSSCh.110...74M"><span>New Defective Brannerite-Type Vanadates. I. Synthesis and Study of Mn 1- x- yφ x<span class="hlt">Na</span> yV <span class="hlt">2-2</span> x-yMo <span class="hlt">2</span> x+yO 6 Solid Solutions</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Masłowska, Bogna; Ziółkowski, Jacek</p> <p>1994-05-01</p> <p>MnV <span class="hlt">2</span>O 6 of the brannerite-type structure (below 540°C) doped with MoO 3 and <span class="hlt">Na</span> <span class="hlt">2</span>O forms isomorphous solid solutions Mn<span class="hlt">Na</span>φ = Mn 1- x-yφ x<span class="hlt">Na</span> yV <span class="hlt">2-2</span> x-yMo <span class="hlt">2</span> x+ yO 6 (φ cation vacancy in the original Mn position), belonging to the pseudoternary MnV <span class="hlt">2</span>O 6-<span class="hlt">Na</span>VMoO 6-MoO 3 system. Particular cases are Mn<span class="hlt">Na</span> = Mn 1- y<span class="hlt">Na</span> y V <span class="hlt">2</span>- yMo yO 6 ( x = 0), Mnφ = Mn 1- xφ xV <span class="hlt">2-2</span> xMo <span class="hlt">2</span> xO 6 ( y = 0), and <span class="hlt">Na</span>φ = <span class="hlt">Na</span> 1- xφ xV 1- xMo 1+ xO 6 ( x + y = 1). MnV <span class="hlt">2</span>O 6 and <span class="hlt">Na</span>VMoO 6 show miscibility in the entire composition range (Mn<span class="hlt">Na</span>). The opposite boundary of Mn<span class="hlt">Na</span>φ passes through the (100 x, 100 y) points (45, 0), (33, 30), and (30, 70). The phase diagram of the pseudobinary MnV <span class="hlt">2</span>O 6-<span class="hlt">Na</span>VMoO 6 system (determined with DTA) shows (i) a narrow double-lens-type solidus-liquidus gap at high values of y , (ii) two peritectic meltings at lower y (yielding the high temperature β-Mn<span class="hlt">Na</span> and Mn <span class="hlt">2</span>V <span class="hlt">2</span>O 7), and (iii) little area of β-Mn<span class="hlt">Na</span>. Lattice parameters of Mn<span class="hlt">Na</span> (determined with X-ray diffraction) reveal small deviations from Vegard's law. As the ionic radii of both dopants (<span class="hlt">Na</span> + and Mo 6+) are, respectively, larger than those of mother ions (Mn <span class="hlt">2</span>+ and V 5+), the unit cell increases in all directions with rising y along the Mn<span class="hlt">Na</span> series of solid solutions. However, due to the anisotropy of the structure, parameter c is strongly sensitive to <span class="hlt">Na</span>/Mn substitution, b is ruled by Mo/V, and a is weakly influenced by Mo/V. Close analogy to the behavior of the previously studied MnV <span class="hlt">2</span>O 6-LiVMoO 6-MoO 6 system is discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28936500','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28936500"><span>Confined <span class="hlt">Na</span>AlH4 nanoparticles inside CeO<span class="hlt">2</span> hollow nanotubes towards enhanced hydrogen storage.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Gao, Qili; Xia, Guanglin; Yu, Xuebin</p> <p>2017-10-05</p> <p><span class="hlt">Na</span>AlH 4 has been widely regarded as a potential hydrogen storage material due to its favorable thermodynamics and high energy density. The high activation energy barrier and high dehydrogenation temperature, however, significantly hinder its practical application. In this paper, CeO <span class="hlt">2</span> hollow nanotubes (HNTs) prepared by a simple electrospinning technique are adopted as functional scaffolds to support <span class="hlt">Na</span>AlH 4 nanoparticles (NPs) towards advanced hydrogen storage performance. The nanoconfined <span class="hlt">Na</span>AlH 4 inside CeO <span class="hlt">2</span> HNTs, synthesized via the infiltration of molten <span class="hlt">Na</span>AlH 4 into the CeO <span class="hlt">2</span> HNTs under high hydrogen pressure, exhibited significantly improved dehydrogenation properties compared with both bulk and ball-milled CeO <span class="hlt">2</span> HNTs-catalyzed <span class="hlt">Na</span>AlH 4 . The onset dehydrogenation temperature of the <span class="hlt">Na</span>AlH 4 @CeO <span class="hlt">2</span> composite was reduced to below 100 °C, with only one main dehydrogenation peak appearing at 130 °C, which is 120 °C and 50 °C lower than for its bulk counterpart and for the ball-milled CeO <span class="hlt">2</span> HNTs-catalyzed <span class="hlt">Na</span>AlH 4 , respectively. Moreover, ∼5.09 wt% hydrogen could be released within 30 min at 180 °C, while only 1.6 wt% hydrogen was desorbed from the ball-milled <span class="hlt">Na</span>AlH 4 under the same conditions. This significant improvement is mainly attributed to the synergistic effects contributed by the CeO <span class="hlt">2</span> HNTs, which could act as not only a structural scaffold to fabricate and confine the <span class="hlt">Na</span>AlH 4 NPs, but also as an effective catalyst to enhance the hydrogen storage performance of <span class="hlt">Na</span>AlH 4 .</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22475638-structural-optical-magnetic-properties-na-sub-eu-sub-si-sub-sub-sub-na-sub-eu-sub-ge-sub-sub-sub-europium-ii-quaternary-chalcogenides-contain-ethane-like-si-sub-sub-sup-ge-sub-sub-sup-moiety','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22475638-structural-optical-magnetic-properties-na-sub-eu-sub-si-sub-sub-sub-na-sub-eu-sub-ge-sub-sub-sub-europium-ii-quaternary-chalcogenides-contain-ethane-like-si-sub-sub-sup-ge-sub-sub-sup-moiety"><span>Structural, optical, and magnetic properties of <span class="hlt">Na</span>{sub 8}Eu{sub <span class="hlt">2</span>}(Si{sub <span class="hlt">2</span>}S{sub 6}){sub <span class="hlt">2</span>} and <span class="hlt">Na</span>{sub 8}Eu{sub <span class="hlt">2</span>}(Ge{sub <span class="hlt">2</span>}S{sub 6}){sub <span class="hlt">2</span>}: Europium(II) quaternary chalcogenides that contain an ethane-like (Si{sub <span class="hlt">2</span>}S{sub 6}){sup 6−} or (Ge{sub <span class="hlt">2</span>}S{sub 6}){sup 6−} moiety</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>Choudhury, Amitava, E-mail: choudhurya@mst.edu; Ghosh, Kartik; Grandjean, Fernande</p> <p>2015-03-15</p> <p>Two isostructural europium(II) quaternary chalcogenides, <span class="hlt">Na</span>{sub 8}Eu{sub <span class="hlt">2</span>}(Si{sub <span class="hlt">2</span>}S{sub 6}){sub <span class="hlt">2</span>}, 1, and <span class="hlt">Na</span>{sub 8}Eu{sub <span class="hlt">2</span>}(Ge{sub <span class="hlt">2</span>}S{sub 6}){sub <span class="hlt">2</span>}, <span class="hlt">2</span>, containing an ethane-like (Si{sub <span class="hlt">2</span>}S{sub 6}){sup 6−} or (Ge{sub <span class="hlt">2</span>}S{sub 6}){sup 6−} moiety have been synthesized by employing the polychalcogenide molten flux method. Single-crystal X-ray diffraction reveals that both compounds crystallize in the C<span class="hlt">2</span>/m space group, and their structures contain layers of ([<span class="hlt">Na</span>{sub <span class="hlt">2</span>}Eu{sub <span class="hlt">2</span>}(Si{sub <span class="hlt">2</span>}S{sub 6}){sub <span class="hlt">2</span>}]{sup 6−}){sub ∞} or ([<span class="hlt">Na</span>{sub <span class="hlt">2</span>}Eu{sub <span class="hlt">2</span>}(Ge{sub <span class="hlt">2</span>}S{sub 6}){sub <span class="hlt">2</span>}]{sup 6−}){sub ∞} anions held together by six interlayer sodium cations to yield (<span class="hlt">Na</span>{sub 6}[<span class="hlt">Na</span>{sub <span class="hlt">2</span>}Eu{sub <span class="hlt">2</span>}(Si{sub <span class="hlt">2</span>}S{sub 6}){sub <span class="hlt">2</span>}]){sub ∞} and (<span class="hlt">Na</span>{sub 6}[<span class="hlt">Na</span>{submore » <span class="hlt">2</span>}Eu{sub <span class="hlt">2</span>}(Ge{sub <span class="hlt">2</span>}S{sub 6}){sub <span class="hlt">2</span>}]){sub ∞}. Compound <span class="hlt">2</span> is a semiconductor with an optical band gap of <span class="hlt">2.15(2</span>) eV. The temperature dependence of the magnetic susceptibility indicates that compounds 1 and <span class="hlt">2</span> are paramagnetic with μ{sub eff}=7.794(1) μ{sub B} per Eu and g=1.964(1) for 1 and μ{sub eff}=8.016(1) μ{sub B} per Eu and g=<span class="hlt">2</span>.020(1) for <span class="hlt">2</span>, moments that are in good agreement with the europium(II) spin-only moment of 7.94 μ{sub B}. The europium-151 Mössbauer isomer shift of <span class="hlt">2</span> confirms the presence of europium(II) cations with an electronic configuration between [Xe]4f{sup 6.81} and 4f{sup 7}6s{sup 0.32}. - Graphical abstract: TOC figure caption: structure of <span class="hlt">Na</span>{sub 8}Eu{sub <span class="hlt">2</span>}(Si{sub <span class="hlt">2</span>}S{sub 6}){sub <span class="hlt">2</span>} viewed along the a-axis showing the filling of A–B and B–A types of anion layers with two different types of cations. - Highlights: • Synthesis of quaternary europium chalcogenides containing ethane-like dimer. • Structural characterization employing single-crystal X-ray diffraction. • Mössbauer spectroscopy and magnetic measurements confirm presence of Eu(II)« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2802203','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2802203"><span><span class="hlt">Na</span>+ channel regulation by Ca<span class="hlt">2</span>+/calmodulin and Ca<span class="hlt">2</span>+/calmodulin-dependent protein kinase II in guinea-pig ventricular myocytes†</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Aiba, Takeshi; Hesketh, Geoffrey G.; Liu, Ting; Carlisle, Rachael; Villa-Abrille, Maria Celeste; O'Rourke, Brian; Akar, Fadi G.; Tomaselli, Gordon F.</p> <p>2010-01-01</p> <p>Aims Calmodulin (CaM) regulates <span class="hlt">Na</span>+ channel gating through binding to an IQ-like motif in the C-terminus. Ca<span class="hlt">2</span>+/CaM-dependent protein kinase II (CaMKII) regulates Ca<span class="hlt">2</span>+ handling, and chronic overactivity of CaMKII is associated with left ventricular hypertrophy and dysfunction and lethal arrhythmias. However, the acute effects of Ca<span class="hlt">2</span>+/CaM and CaMKII on cardiac <span class="hlt">Na</span>+ channels are not fully understood. Methods and results Purified <span class="hlt">Na</span>V1.5–glutathione-S-transferase fusion peptides were phosphorylated in vitro by CaMKII predominantly on the I–II linker. Whole-cell voltage-clamp was used to measure <span class="hlt">Na</span>+ current (INa) in isolated guinea-pig ventricular myocytes in the absence or presence of CaM or CaMKII in the pipette solution. CaMKII shifted the voltage dependence of <span class="hlt">Na</span>+ channel availability by ≈+5 mV, hastened recovery from inactivation, decreased entry into intermediate or slow inactivation, and increased persistent (late) current, but did not change INa decay. These CaMKII-induced changes of <span class="hlt">Na</span>+ channel gating were completely abolished by a specific CaMKII inhibitor, autocamtide-<span class="hlt">2</span>-related inhibitory peptide (AIP). Ca<span class="hlt">2</span>+/CaM alone reproduced the CaMKII-induced changes of INa availability and the fraction of channels undergoing slow inactivation, but did not alter recovery from inactivation or the magnitude of the late current. Furthermore, the CaM-induced changes were also completely abolished by AIP. On the other hand, cAMP-dependent protein kinase A inhibitors did not abolish the CaM/CaMKII-induced alterations of INa function. Conclusion Ca<span class="hlt">2</span>+/CaM and CaMKII have distinct effects on the inactivation phenotype of cardiac <span class="hlt">Na</span>+ channels. The differences are consistent with CaM-independent effects of CaMKII on cardiac <span class="hlt">Na</span>+ channel gating. PMID:19797425</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016amsf.conf..335Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016amsf.conf..335Y"><span>Effect of <span class="hlt">Na</span><span class="hlt">2</span>O on Crystallisation Behaviour and Heat Transfer of Fluorine-Free Mould Fluxes</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yang, Jian; Zhang, Jianqiang; Sasaki, Yasushi; Ostrovski, Oleg; Zhang, Chen; Cai, Dexiang; Kashiwaya, Yoshiaki</p> <p></p> <p>Most of the commercial mould fluxes contain fluorides which bring about serious environmental problems. The major challenge in the application of fluorine-free mould fluxes is to control the heat transfer from the strand to copper mould which is closely related to crystallisation behaviour. In this study, the effects of <span class="hlt">Na</span><span class="hlt">2</span>O on the crystallisation behaviour and heat transfer of CaO-SiO<span class="hlt">2</span>-<span class="hlt">Na</span><span class="hlt">2</span>O-B<span class="hlt">2</span>O3-TiO<span class="hlt">2</span>-Al<span class="hlt">2</span>O3-MgO-Li<span class="hlt">2</span>O mould fluxes were investigated using single /double hot thermocouple technique (SHTT/DHTT) and infrared emitter technique (IET), respectively. Continuous cooling transformation (CCT) and time-temperature transformation (TTT) diagrams constructed using SHTT showed that the increase of <span class="hlt">Na</span><span class="hlt">2</span>O concentration led to higher critical cooling rate and shorter incubation time. The crystallisation behaviour in a thermal gradient was examined using DHTT. The heat flux measured by IET showed that the increase of <span class="hlt">Na</span><span class="hlt">2</span>O concentration decreased the heat flux when <span class="hlt">Na</span><span class="hlt">2</span>O was lower than 9 mass% but the further increase of <span class="hlt">Na</span><span class="hlt">2</span>O raised the heat flux. The relationship between flux crystallisation and heat transfer was also discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018AIPC.1942m0020B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AIPC.1942m0020B"><span>Magnetic ground state of the layered honeycomb compound <span class="hlt">Na</span><span class="hlt">2</span>Co<span class="hlt">2</span>TeO6</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bera, A. K.; Yusuf, S. M.</p> <p>2018-04-01</p> <p>The magnetic correlations in the <span class="hlt">2</span>D layered honeycomb compound <span class="hlt">Na</span><span class="hlt">2</span>Co<span class="hlt">2</span>TeO6 has been investigated. The temperature dependent susceptibility curve reveals a transition to the magnetically ordered state at TN ˜ 25 K. The temperature dependent neutron diffraction study confirms an antiferromagnetic ordering below TN. The magnetic ground state is determined to be a zigzag antiferromagnet that appears due to competing exchange interactions beyond nearest neighbors. The moments align along the crystallographic b axis with reduced ordered magnetic moment values of <span class="hlt">2.72(2</span>) μB/Co<span class="hlt">2</span>+ and <span class="hlt">2</span>.52(3) μB/Co<span class="hlt">2</span>+ for two Co sites, respectively. In comparison to the theoretical phase diagram the determined zigzag antiferromagnetic ground state suggests that the compound <span class="hlt">Na</span><span class="hlt">2</span>Co<span class="hlt">2</span>TeO6 is situated in the proximity to the quantum spin liquid state in the phase diagram.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AIPC.1903e0013S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AIPC.1903e0013S"><span>Effect of <span class="hlt">Na</span><span class="hlt">2</span>SiO3/<span class="hlt">Na</span>OH on mechanical properties and microstructure of geopolymer mortar using fly ash and rice husk ash as precursor</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Saloma, Hanafiah, Elysandi, Debby Orjina; Meykan, Della Garnesia</p> <p>2017-11-01</p> <p>Geopolymer concrete is an eco-friendly concrete that can reduce carbon emissions on the earth surface because it used industrial waste material such as fly ash, rice husk ash, bagasse ash, and palm oil fuel. Geopolymer is semi-crystalline amorphous materials which has irregular chemical bonds structure. The material is produced by geosynthesis of aluminosilicates and alkali-silicates which produce the Si-O-Al polymer structure. This research used the ratio of fly ash and rice husk ash as precursors e.g. 100:0, 75:25, 50:50, and 25:75. <span class="hlt">Na</span>OH solutions of 14 M and <span class="hlt">Na</span><span class="hlt">2</span>SiO3 solutions with the variation e.g. <span class="hlt">2</span>.5, <span class="hlt">2</span>.75, 3.00, and 3.25 were used as activators on mortar geopolymer mixture. The tests of fresh mortar were slump flow and setting time. The optimum compressive strength is 68.36 MPa for 28 days resulted from mixture using 100% fly ash and <span class="hlt">Na</span><span class="hlt">2</span>SiO3 and <span class="hlt">Na</span>OH with ratio <span class="hlt">2</span>.75. The largest value of slump flow test resulted from mixture using <span class="hlt">Na</span><span class="hlt">2</span>SiO3 and <span class="hlt">Na</span>OH with ratio <span class="hlt">2</span>.50 is 17.25 cm. Based on SEM test results, mortar geopolymer microstructure with mixture RHA 0% has less pores and denser CSH structure.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017PhRvP...7f4003Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017PhRvP...7f4003Z"><span>Effects of Transition-Metal Mixing on <span class="hlt">Na</span> Ordering and Kinetics in Layered P <span class="hlt">2</span> Oxides</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zheng, Chen; Radhakrishnan, Balachandran; Chu, Iek-Heng; Wang, Zhenbin; Ong, Shyue Ping</p> <p>2017-06-01</p> <p>Layered P <span class="hlt">2</span> oxides are promising cathode materials for rechargeable sodium-ion batteries. In this work, we systematically investigate the effects of transition-metal (TM) mixing on <span class="hlt">Na</span> ordering and kinetics in the NaxCo1 -yMnyO<span class="hlt">2</span> model system using density-functional-theory (DFT) calculations. The DFT-predicted 0-K stability diagrams indicate that Co-Mn mixing reduces the energetic differences between <span class="hlt">Na</span> orderings, which may account for the reduction of the number of phase transformations observed during the cycling of mixed-TM P <span class="hlt">2</span> layered oxides compared to a single TM. Using ab initio molecular-dynamics simulations and nudged elastic-band calculations, we show that the TM composition at the <span class="hlt">Na</span>(1) (face-sharing) site has a strong influence on the <span class="hlt">Na</span> site energies, which in turn impacts the kinetics of <span class="hlt">Na</span> diffusion towards the end of the charge. By employing a site-percolation model, we establish theoretical upper and lower bounds for TM concentrations based on their effect on <span class="hlt">Na</span>(1) site energies, providing a framework to rationally tune mixed-TM compositions for optimal <span class="hlt">Na</span> diffusion.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017INL.....7..210N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017INL.....7..210N"><span>Novel synthesis approach for stable sodium superoxide (<span class="hlt">Na</span>O<span class="hlt">2</span>) nanoparticles for LPG sensing application</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Nemade, Kailash; Waghuley, Sandeep</p> <p>2017-05-01</p> <p>The synthesis of stable superoxide is still great challenge for the researchers working in the field of materials science. Through this letter, we report the novel and simple synthesis approach for the preparation of stable sodium superoxide (<span class="hlt">Na</span>O<span class="hlt">2</span>) nanoparticles. <span class="hlt">Na</span>O<span class="hlt">2</span> nanoparticles were prepared by a spray pyrolysis technique, under oxygen rich environment for gas sensing application. The texture characterizations show that as-obtained <span class="hlt">Na</span>O<span class="hlt">2</span> nanoparticles have high structural purity. Most importantly, <span class="hlt">Na</span>O<span class="hlt">2</span> nanoparticles exhibits higher sensing response, shorter response time and recovery time, low operating temperature and good stability during sensing of liquefied petroleum gas (LPG). The main accomplishment of present work is that as-fabricated sensor has low operating temperature (423 K), which is below auto-ignition temperature of LPG. The gas sensing mechanism of <span class="hlt">Na</span>O<span class="hlt">2</span> nanoparticles was discussed without the conventional oxygen bridging mechanism. Through this short communication, LPG sensing application of stable sodium superoxide nanoparticle is explored.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70016658','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70016658"><span>On the entropy of glaucophane <span class="hlt">Na</span><span class="hlt">2</span>Mg3Al<span class="hlt">2</span>Si8O22(OH)<span class="hlt">2</span></span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Robie, R.A.; Hemingway, B.S.; Gillet, P.; Reynard, B.</p> <p>1991-01-01</p> <p>The heat capacity of glaucophane from the Sesia-Lanza region of Italy having the approximate composition (<span class="hlt">Na</span>1.93Ca0.05Fe0.02) (Mg<span class="hlt">2</span>.60Fe0.41) (Al1.83Fe0.15Cr0.01) (Si7.92Al0.08)O22(OH)<span class="hlt">2</span> was measured by adiabatic calorimetry between 4.6 and 359.4 K. After correcting the Cp0data to values for ideal glaucophane, <span class="hlt">Na</span><span class="hlt">2</span>Mg3Al<span class="hlt">2</span>Si8O22(OH)<span class="hlt">2</span> the third-law entropy S2980-S00was calculated to be 541.<span class="hlt">2</span>??3.0 J??mol-1??K-1. Our value for S2980-S00is 12.0 J??mol-1??K-1 (<span class="hlt">2.2</span>%) smaller than the value of Likhoydov et al. (1982), 553.<span class="hlt">2</span>??3.0, is within 6.<span class="hlt">2</span> J??mol-1??K-1 of the value estimated by Holland (1988), and agrees remarkably well with the value calculated by Gillet et al. (1989) from spectroscopic data, 539 J??mol-1??K-1. ?? 1991 Springer-Verlag.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19940011718','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19940011718"><span>Grosnaja ABCs: Magnesium isotope compositions</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Goswami, J. N.; Srinivasan, G.; Ulyanov, A. A.</p> <p>1993-01-01</p> <p>Three <span class="hlt">CAI</span>'s from the Grosnaja CV3 chondrite were analyzed for their magnesium isotopic compositions by the ion microprobe. The selected <span class="hlt">CAI</span>'s represent three distinct types: GR4(compact Type A), GR7(Type B) and GR<span class="hlt">2</span>(Type C). Petrographic studies indicate that all three Grosnaja inclusions were subjected to secondary alterations. The Type A <span class="hlt">CAI</span> GR4 is primarily composed of melilite with spinel and pyroxene occurring as minor phases. The rim of the inclusion does not exhibit distinct layered structure and secondary alteration products (garnet, Fe-rich olivine and <span class="hlt">Na</span>-rich plagioclase) are present in some localized areas near the rim region. The average major element compositions of different mineral phases in GR4 are given. Preliminary REE data suggest a depletion of HREE relative to LREE by about a factor of 3 without any clear indication of interelement fractionation. The <span class="hlt">CAI</span> GR7 has textural and minerological characteristics similar to Type B inclusions. The REE data show a pattern that is similar to Group 6 with enrichment in Eu and Yb. In addition, a depletion of HREE compared to LREE is also evident in this object. Melilite composition shows a broad range of akermanite content (Ak(sub 15-55)). Detailed petrographic study is in progress. GR<span class="hlt">2</span> is a anorthite-rich Type C inclusion with large plagioclase laths intergrown with Ti-rich pyroxene. The average plagioclase composition is close to pure anorthite (An99).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25129762','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25129762"><span>Removing polysaccharides-and saccharides-related coloring impurities in alkyl polyglycosides by bleaching with the H<span class="hlt">2</span>O<span class="hlt">2</span>/TAED/<span class="hlt">Na</span>HCO3 system.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Yanmei, Liu; Jinliang, Tao; Jiao, Sun; Wenyi, Chen</p> <p>2014-11-04</p> <p>The effect of H<span class="hlt">2</span>O<span class="hlt">2</span>/TAED/<span class="hlt">Na</span>HCO3 system, namely <span class="hlt">Na</span>HCO3 as alkaline agent with the (tetra acetyl ethylene diamine (TAED)) TAED-activated peroxide system, bleaching of alkyl polyglycosides solution was studied by spectrophotometry. The results showed that the optimal bleaching conditions about H<span class="hlt">2</span>O<span class="hlt">2</span>/TAED/<span class="hlt">Na</span>HCO3 system bleaching of alkyl polyglycosides solution were as follows: molar ratio of TAED to H<span class="hlt">2</span>O<span class="hlt">2</span> was 0.06, addition of H<span class="hlt">2</span>O<span class="hlt">2</span> was 8.6%, addition of <span class="hlt">Na</span>HCO3 was 3.<span class="hlt">2</span>%, bleaching temperature of 50-65 °C, addition of MgO was 0.13%, and bleaching time was 8h. If too much amount of <span class="hlt">Na</span>HCO3 was added to the system and maintained alkaline pH, the bleaching effect would be greatly reduced. Fixing molar ratio of TAED to H<span class="hlt">2</span>O<span class="hlt">2</span> and increasing the amount of H<span class="hlt">2</span>O<span class="hlt">2</span> were beneficial to improve the whiteness of alkyl polyglycosides, but adding too much amount of H<span class="hlt">2</span>O<span class="hlt">2</span> would reduce the transparency. In the TAED-activated peroxide system, <span class="hlt">Na</span>HCO3 as alkaline agent and buffer agent, could overcome the disadvantage of producing black precipitates when <span class="hlt">Na</span>OH as alkaline agent. Copyright © 2014 Elsevier Ltd. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/18601978','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/18601978"><span>Reduced expression of <span class="hlt">Na</span>(v)1.6 sodium channels and compensation by <span class="hlt">Na</span>(v)1.<span class="hlt">2</span> channels in mice heterozygous for a null mutation in Scn8a.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Vega, Ana V; Henry, Diane L; Matthews, Gary</p> <p>2008-09-05</p> <p>The voltage-gated sodium channel alpha subunit <span class="hlt">Na</span>(v)1.6, encoded by the Scn8a gene, accumulates at high density at mature nodes of Ranvier of myelinated axons, replacing the <span class="hlt">Na</span>(v)1.<span class="hlt">2</span> channels found at nodes earlier in development. To investigate this preferential expression of <span class="hlt">Na</span>(v)1.6 at adult nodes, we examined isoform-specific expression of sodium channels in mice heterozygous for a null mutation in Scn8a. Immunoblots from these +/- mice had 50% of the wild-type level of <span class="hlt">Na</span>(v)1.6 protein, and their optic-nerve nodes of Ranvier had correspondingly less anti-<span class="hlt">Na</span>(v)1.6 immunofluorescence. Protein level and nodal immunofluorescence of the <span class="hlt">Na</span>(v)1.<span class="hlt">2</span> alpha subunit increased in Scn8a(+/-) mice, keeping total sodium channel expression approximately constant despite partial loss of <span class="hlt">Na</span>(v)1.6 channels. The results are consistent with a model in which <span class="hlt">Na</span>(v)1.6 and <span class="hlt">Na</span>(v)1.<span class="hlt">2</span> compete for binding partners at sites of high channel density, such as nodes of Ranvier. We suggest that <span class="hlt">Na</span>(v)1.6 channels normally occupy most of the molecular machinery responsible for channel clustering because they have higher binding affinity, and not because they are exclusively recognized by mechanisms for transport and insertion of sodium channels in myelinated axons. The reduced amount of <span class="hlt">Na</span>(v)1.6 protein in Scn8a(+/-) mice is apparently insufficient to saturate the nodal binding sites, allowing <span class="hlt">Na</span>(v)1.<span class="hlt">2</span> channels to compete more successfully.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5466374','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5466374"><span>MnTiO3-driven low-temperature oxidative coupling of methane over TiO<span class="hlt">2</span>-doped Mn<span class="hlt">2</span>O3-<span class="hlt">Na</span><span class="hlt">2</span>WO4/SiO<span class="hlt">2</span> catalyst</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Wang, Pengwei; Zhao, Guofeng; Wang, Yu; Lu, Yong</p> <p>2017-01-01</p> <p>Oxidative coupling of methane (OCM) is a promising method for the direct conversion of methane to ethene and ethane (C<span class="hlt">2</span> products). Among the catalysts reported previously, Mn<span class="hlt">2</span>O3-<span class="hlt">Na</span><span class="hlt">2</span>WO4/SiO<span class="hlt">2</span> showed the highest conversion and selectivity, but only at 800° to 900°C, which represents a substantial challenge for commercialization. We report a TiO<span class="hlt">2</span>-doped Mn<span class="hlt">2</span>O3-<span class="hlt">Na</span><span class="hlt">2</span>WO4/SiO<span class="hlt">2</span> catalyst by using Ti-MWW zeolite as TiO<span class="hlt">2</span> dopant as well as SiO<span class="hlt">2</span> support, enabling OCM with 26% conversion and 76% C<span class="hlt">2</span>-C3 selectivity at 720°C because of MnTiO3 formation. MnTiO3 triggers the low-temperature Mn<span class="hlt">2</span>+↔Mn3+ cycle for O<span class="hlt">2</span> activation while working synergistically with <span class="hlt">Na</span><span class="hlt">2</span>WO4 to selectively convert methane to C<span class="hlt">2</span>-C3. We also prepared a practical Mn<span class="hlt">2</span>O3-TiO<span class="hlt">2</span>-<span class="hlt">Na</span><span class="hlt">2</span>WO4/SiO<span class="hlt">2</span> catalyst in a ball mill. This catalyst can be transformed in situ into MnTiO3-<span class="hlt">Na</span><span class="hlt">2</span>WO4/SiO<span class="hlt">2</span>, yielding 22% conversion and 62% selectivity at 650°C. Our results will stimulate attempts to understand more fully the chemistry of MnTiO3-governed low-temperature activity, which might lead to commercial exploitation of a low-temperature OCM process. PMID:28630917</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018ApSS..444..436P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018ApSS..444..436P"><span>Cu(In,Ga)Se<span class="hlt">2</span> surface treatment with <span class="hlt">Na</span> and <span class="hlt">Na</span>F: A combined photoelectron spectroscopy and surface photovoltage study in ultra-high vacuum</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Parvan, V.; Mizrak, A.; Majumdar, I.; Ümsür, B.; Calvet, W.; Greiner, D.; Kaufmann, C. A.; Dittrich, T.; Avancini, E.; Lauermann, I.</p> <p>2018-06-01</p> <p>Either metallic <span class="hlt">Na</span> or <span class="hlt">Na</span>F were deposited onto Cu(In,Ga)Se<span class="hlt">2</span> surfaces and studied by photoelectron spectroscopy and surface photovoltage spectroscopy without breaking the ultra-high vacuum. The deposition of elemental <span class="hlt">Na</span> at room temperature led to the formation of an intermediate Cu and Ga rich layer at the CIGSe surface, whereas for <span class="hlt">Na</span>F the composition of the CIGSe surface remained unchanged. A metal like surface induced by an inverted near surface region with a reduced number of defect states was formed after the deposition of <span class="hlt">Na</span>. Under the chosen experimental conditions, the near surface layer was independent on the amount of <span class="hlt">Na</span> and stable in time. In contrast, the usage of <span class="hlt">Na</span>F weakened the inversion and led to an increased band bending compared to the untreated CIGSe sample. The SPV signals decreased with proceeding time after the deposition of <span class="hlt">Na</span>F.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26391815','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26391815"><span>Extraction of Mg(OH)<span class="hlt">2</span> from Mg silicate minerals with <span class="hlt">Na</span>OH assisted with H<span class="hlt">2</span>O: implications for CO<span class="hlt">2</span> capture from exhaust flue gas.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Madeddu, Silvia; Priestnall, Michael; Godoy, Erik; Kumar, R Vasant; Raymahasay, Sugat; Evans, Michael; Wang, Ruofan; Manenye, Seabelo; Kinoshita, Hajime</p> <p>2015-01-01</p> <p>The utilisation of Mg(OH)<span class="hlt">2</span> to capture exhaust CO<span class="hlt">2</span> has been hindered by the limited availability of brucite, the Mg(OH)<span class="hlt">2</span> mineral in natural deposits. Our previous study demonstrated that Mg(OH)<span class="hlt">2</span> can be obtained from dunite, an ultramafic rock composed of Mg silicate minerals, in highly concentrated <span class="hlt">Na</span>OH aqueous systems. However, the large quantity of <span class="hlt">Na</span>OH consumed was considered an obstacle for the implementation of the technology. In the present study, Mg(OH)<span class="hlt">2</span> was extracted from dunite reacted in solid systems with <span class="hlt">Na</span>OH assisted with H<span class="hlt">2</span>O. The consumption of <span class="hlt">Na</span>OH was reduced by 97% with respect to the <span class="hlt">Na</span>OH aqueous systems, maintaining a comparable yield of Mg(OH)<span class="hlt">2</span> extraction, i.e. 64.8-66%. The capture of CO<span class="hlt">2</span> from a CO<span class="hlt">2</span>-N<span class="hlt">2</span> gas mixture was tested at ambient conditions using a Mg(OH)<span class="hlt">2</span> aqueous slurry. Mg(OH)<span class="hlt">2</span> almost fully dissolved and reacted with dissolved CO<span class="hlt">2</span> by forming Mg(HCO3)<span class="hlt">2</span> which remained in equilibrium storing the CO<span class="hlt">2</span> in the aqueous solution. The CO<span class="hlt">2</span> balance of the process was assessed from the emissions derived from the power consumption for <span class="hlt">Na</span>OH production and Mg(OH)<span class="hlt">2</span> extraction together with the CO<span class="hlt">2</span> captured by Mg(OH)<span class="hlt">2</span> derived from dunite. The process resulted as carbon neutral when dunite is reacted at 250 °C for durations of 1 and 3 hours and CO<span class="hlt">2</span> is captured as Mg(HCO3)<span class="hlt">2</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22645383-crystal-structure-new-polar-borate-na-sub-ce-sub-bo-sub-oh-bo-sub-sub-sub-isolated-boron-triangles','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22645383-crystal-structure-new-polar-borate-na-sub-ce-sub-bo-sub-oh-bo-sub-sub-sub-isolated-boron-triangles"><span>Crystal structure of a new polar borate <span class="hlt">Na</span>{sub <span class="hlt">2</span>}Ce{sub <span class="hlt">2</span>}[BO{sub <span class="hlt">2</span>}(OH)][BO{sub 3}]{sub <span class="hlt">2</span>} · H{sub <span class="hlt">2</span>}O with isolated boron triangles</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>Topnikova, A. P.; Belokoneva, E. L., E-mail: elbel@geol.msu.ru; Dimitrova, O. V.</p> <p>2016-11-15</p> <p>Crystals of a new polar borate <span class="hlt">Na</span>{sub <span class="hlt">2</span>}Ce{sub <span class="hlt">2</span>}[BO{sub <span class="hlt">2</span>}(OH)][BO{sub 3}]{sub <span class="hlt">2</span>} · H{sub <span class="hlt">2</span>}O were prepared by hydrothermal synthesis. The crystals are orthorhombic, a = 7.2295(7) Å, b = 11.2523(8) Å, c = 5.1285(6) Å, Z = <span class="hlt">2</span>, sp. gr. C<span class="hlt">2</span>mm (Amm<span class="hlt">2</span>), R = 0.0253. The formula of the compound was derived from the structure determination. The Ce and <span class="hlt">Na</span> atoms are coordinated by nine and six O atoms, respectively. The Ce position is split, and a small amount of Ce is incorporated into the <span class="hlt">Na</span>1 site with the isomorphous substitution for <span class="hlt">Na</span>. The anionic moieties exist as isolatedmore » BO{sub 3} and BO{sub <span class="hlt">2</span>}(OH) triangles. The planes of the BO{sub <span class="hlt">2</span>}(OH) triangles with mm<span class="hlt">2</span> symmetry are parallel to the ab plane. The planes of the BO{sub 3} triangles with m symmetry are perpendicular to the ab plane and are rotated in a diagonal way. The splitting of the Ce positions and the polar arrangement of the BO{sub <span class="hlt">2</span>}(OH) triangles, water molecules, and <span class="hlt">Na</span> atoms are observed along the polar a axis. The new structure is most similar to the new borate <span class="hlt">Na</span>Ca{sub 4}[BO{sub 3}]{sub 3} (sp. gr. Ama<span class="hlt">2</span>), in which triangles of one type are arranged in a polar fashion along the c axis. Weak nonlinear-optical properties of both polar borates are attributed to the quenching of the second-harmonic generation due to the mutually opposite orientation of two-thirds of B triangles in the unit cell.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22341707-luminescent-properties-na-sub-casio-sub-eu-sup-its-potential-application-white-light-emitting-diodes','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22341707-luminescent-properties-na-sub-casio-sub-eu-sup-its-potential-application-white-light-emitting-diodes"><span>Luminescent properties of <span class="hlt">Na</span>{sub <span class="hlt">2</span>}CaSiO{sub 4}:Eu{sup <span class="hlt">2</span>+} and its potential application in white light emitting diodes</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Wang, Zhijun, E-mail: wangzhijunmail@yahoo.com.cn; Li, Panlai; Li, Ting</p> <p>2013-06-01</p> <p>Graphical abstract: <span class="hlt">Na</span>{sub <span class="hlt">2</span>}CaSiO{sub 4}:Eu{sup <span class="hlt">2</span>+} phosphor can be effectively excited by an ultraviolet and near-ultraviolet light, and produce a bright blue emission centered at 436 nm. The CIE chromaticity coordinations (x, y) of <span class="hlt">Na</span>{sub <span class="hlt">2</span>}CaSiO{sub 4}:Eu{sup <span class="hlt">2</span>+}(NSCE)/Li{sub <span class="hlt">2</span>}SrSiO{sub 4}:Eu{sup <span class="hlt">2</span>+}(LSSE) vary with the molar ratio of the two constituents. When NSCE/LSSE is 1:3, the CIE chromaticity coordination is (0.332, 0.346), which is close to that of the natural sunlight (0.33, 0.33). The results indicate that <span class="hlt">Na</span>{sub <span class="hlt">2</span>}CaSiO{sub 4}:Eu{sup <span class="hlt">2</span>+} may be a promising blue phosphor for UV chip-based multi-phosphor converted white light emitting diodes. Highlights: ► <span class="hlt">Na</span>{sub <span class="hlt">2</span>}CaSiO{sub 4}:Eu{supmore » <span class="hlt">2</span>+} shows the blue emission with a peak at 436 nm and broad excitation band in the UV/n-UV range. ► White light with CIE coordinates (0.332, 0.346) is generated by mixing the blue phosphor with the Li{sub <span class="hlt">2</span>}SrSiO{sub 4}:Eu{sup <span class="hlt">2</span>+} yellow phosphor. ► <span class="hlt">Na</span>{sub <span class="hlt">2</span>}CaSiO{sub 4}:Eu{sup <span class="hlt">2</span>+} would be a promising blue phosphor candidate for UV chip-based multi-phosphor converted white LEDs. - Abstract: A novel blue phosphor <span class="hlt">Na</span>{sub <span class="hlt">2</span>}CaSiO{sub 4}:Eu{sup <span class="hlt">2</span>+} is synthesized by a high temperature solid-state reaction, and its luminescent properties are systematically studied. <span class="hlt">Na</span>{sub <span class="hlt">2</span>}CaSiO{sub 4}:Eu{sup <span class="hlt">2</span>+} can be effectively excited by the 354 nm radiation, and create blue emission (436 nm). The emission intensity of <span class="hlt">Na</span>{sub <span class="hlt">2</span>}CaSiO{sub 4}:Eu{sup <span class="hlt">2</span>+} is influenced by the Eu{sup <span class="hlt">2</span>+} doping content, and the optimal doping content is 1.5%, and the concentration quenching mechanism of Eu{sup <span class="hlt">2</span>+} in <span class="hlt">Na</span>{sub <span class="hlt">2</span>}CaSiO{sub 4} can be attributed to the multipolar interaction. The white light with CIE coordinates (0.332, 0.346) is generated by mixing the blue phosphor <span class="hlt">Na</span>{sub <span class="hlt">2</span>}CaSiO{sub 4}:Eu{sup <span class="hlt">2</span>+} with the yellow phosphor Li{sub <span class="hlt">2</span>}SrSiO{sub 4}:Eu{sup <span class="hlt">2</span>+}. The results indicate that <span class="hlt">Na</span>{sub <span class="hlt">2</span>}CaSiO{sub 4}:Eu{sup <span class="hlt">2</span>+} may be a potential blue emitting phosphor for UV chip-based multi</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li class="active"><span>20</span></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_20 --> <div id="page_21" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li class="active"><span>21</span></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="401"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018ApSS..427.1183Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018ApSS..427.1183Z"><span><span class="hlt">Na</span><span class="hlt">2</span>Ti6O13@TiO<span class="hlt">2</span> core-shell nanorods with controllable mesoporous shells and their enhanced photocatalytic performance</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zhou, Xuefan; Zhong, Donglin; Luo, Hang; Pan, Jun; Zhang, Dou</p> <p>2018-01-01</p> <p>In this study, dispersive and free-standing <span class="hlt">Na</span><span class="hlt">2</span>Ti6O13 nanorods with diameter of about 500 nm and length of about 10 μm were synthesized by the molten salt method. The <span class="hlt">Na</span><span class="hlt">2</span>Ti6O13@TiO<span class="hlt">2</span> (denoted as TTO) core-shell nanorods were fabricated by a versatile kinetics controlled coating method. The TiO<span class="hlt">2</span> shells were uniform and mesoporous with exposed {101} facets. The thickness of TiO<span class="hlt">2</span> shells can be well controlled by the content of Ti(OC4H9)4, ranging from 0 nm, 15 nm, 60 nm to 70 nm corresponding to <span class="hlt">Na</span><span class="hlt">2</span>Ti6O13, 0.25-TTO, 0.50-TTO and 0.75-TTO nanorods respectively. The crystalline phases, microstructure, porosity, photoabsorption and photocatalytic performance of all the samples were investigated systematically. The nanoscale heterojunction structure between <span class="hlt">Na</span><span class="hlt">2</span>Ti6O13 and TiO<span class="hlt">2</span>, reductive TiO<span class="hlt">2</span> {101} facets and high aspect ratio <span class="hlt">Na</span><span class="hlt">2</span>Ti6O13 nanorods resulted in the enhanced photocatalytic performance of TTO nanorods. The optimized thickness of TiO<span class="hlt">2</span> shells were about 60 nm for 0.50-TTO nanorods, which possessed superior BET surface area, optical absorption and photocatalytic performance.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016MMTB...47.2440S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016MMTB...47.2440S"><span>Phase Equilibria Study in the TeO<span class="hlt">2</span>-<span class="hlt">Na</span><span class="hlt">2</span>O-SiO<span class="hlt">2</span> System in Air Between 723 K (500 °C) and 1473 K (1200 °C)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Santoso, Imam; Taskinen, Pekka</p> <p>2016-08-01</p> <p>Knowledge of phase equilibria in the TeO<span class="hlt">2</span>-<span class="hlt">Na</span><span class="hlt">2</span>O-SiO<span class="hlt">2</span> system at elevated temperatures is important for ceramic and glass industries and for improving the operation of the smelting process of tellurium-containing materials. A review of previous investigations has indicated, however, that there are omissions in the available datasets on the liquidus temperatures of the molten TeO<span class="hlt">2</span>-<span class="hlt">Na</span><span class="hlt">2</span>O-SiO<span class="hlt">2</span> mixtures. The employed experimental method included equilibration of mixtures made from high purity oxides, rapid quenching of the equilibrated samples in water and followed by compositional analysis of the phases using an electron probe X-ray microanalyzer. The liquidus and phase equilibria in the TeO<span class="hlt">2</span>-SiO<span class="hlt">2</span>, TeO<span class="hlt">2</span>-<span class="hlt">Na</span><span class="hlt">2</span>O, and SiO<span class="hlt">2</span>-TeO<span class="hlt">2</span>-<span class="hlt">Na</span><span class="hlt">2</span>O systems have been studied for a wide range of compositions between 723 K (500 °C) and 1473 K (1200 °C) at TeO<span class="hlt">2</span>, SiO<span class="hlt">2</span>, and <span class="hlt">Na</span><span class="hlt">2</span>SiO3 saturations. New data have been generated in the SiO<span class="hlt">2</span>-TeO<span class="hlt">2</span>-<span class="hlt">Na</span><span class="hlt">2</span>O system at SiO<span class="hlt">2</span> saturation. The liquidus compositions in the TeO<span class="hlt">2</span>-<span class="hlt">Na</span><span class="hlt">2</span>O system at TeO<span class="hlt">2</span> saturation have been compared with the previous data and an assessed phase diagram.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22149935-studies-solid-solutions-based-layered-honeycomb-ordered-phases-p2-na-sub-sub-teo-sub-co-ni-zn','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22149935-studies-solid-solutions-based-layered-honeycomb-ordered-phases-p2-na-sub-sub-teo-sub-co-ni-zn"><span>Studies on solid solutions based on layered honeycomb-ordered phases P<span class="hlt">2</span>-<span class="hlt">Na</span>{sub <span class="hlt">2</span>}M{sub <span class="hlt">2</span>}TeO{sub 6} (M=Co, Ni, Zn)</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>Berthelot, Romain; Schmidt, Whitney; Sleight, A.W.</p> <p>2012-12-15</p> <p>Three complete solid solutions between the layered phases P<span class="hlt">2</span>-<span class="hlt">Na</span>{sub <span class="hlt">2</span>}M{sub <span class="hlt">2</span>}TeO{sub 6} (M=Co, Ni, Zn) have been prepared by conventional solid state method and investigated through X-ray diffraction, magnetism and optical measurements. All compositions are characterized by a M{sup <span class="hlt">2</span>+}/X{sup 6+} honeycomb ordering within the slabs and crystallize in a hexagonal unit cell. However, a structural transition based on a different stacking is observed as nickel (space group P6{sub 3}/mcm) is substituted by zinc or cobalt (space group P6{sub 3}22). All compositions exhibit a paramagnetic Curie-Weiss behavior at high temperatures; and the magnetic moment values confirm the presence of Ni{supmore » <span class="hlt">2</span>+} and/or Co{sup <span class="hlt">2</span>+} cations. The low-temperature antiferromagnetic order of <span class="hlt">Na</span>{sub <span class="hlt">2</span>}Ni{sub <span class="hlt">2</span>}TeO{sub 6} and <span class="hlt">Na</span>{sub <span class="hlt">2</span>}Co{sub <span class="hlt">2</span>}TeO{sub 6} is suppressed by zinc substitution. The color of the obtained compositions varies from pink, to light green and white when M=Co, Ni, Zn, respectively. - Graphical abstract: The comparison between the structure of <span class="hlt">Na</span>{sub <span class="hlt">2</span>}Ni{sub <span class="hlt">2</span>}TeO{sub 6} (left) and <span class="hlt">Na</span>{sub <span class="hlt">2</span>}M{sub <span class="hlt">2</span>}TeO{sub 6} (M=Co, Zn) (right) evidences the stacking difference with distinct atom sequences along the hexagonal c-axis. Highlights: Black-Right-Pointing-Pointer Solid solutions between lamellar phases <span class="hlt">Na</span>{sub <span class="hlt">2</span>}M{sub <span class="hlt">2</span>}TeO{sub 6} (M=Co, Ni, Zn) are investigated. Black-Right-Pointing-Pointer A M{sup <span class="hlt">2</span>+}/X{sup 6+} honeycomb ordering characterized all the compositions. Black-Right-Pointing-Pointer A structural transition is shown when Ni is replaced by Co or Zn. Black-Right-Pointing-Pointer The low-temperature AFM ordering of <span class="hlt">Na</span>{sub <span class="hlt">2</span>}Ni{sub <span class="hlt">2</span>}TeO{sub 6} and <span class="hlt">Na</span>{sub <span class="hlt">2</span>}Co{sub <span class="hlt">2</span>}TeO{sub 6} is suppressed by zinc substitution. Black-Right-Pointing-Pointer Color changes from pink to light green and white when M=Co, Ni, Zn, respectively.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012JCrGr.355..109M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012JCrGr.355..109M"><span>Condition of Si crystal formation by vaporizing <span class="hlt">Na</span> from <span class="hlt">Na</span>Si</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Morito, Haruhiko; Karahashi, Taiki; Yamane, Hisanori</p> <p>2012-09-01</p> <p><span class="hlt">Na</span>Si was heated at various <span class="hlt">Na</span> vapor pressures (p<span class="hlt">Na</span> 0.1-1.<span class="hlt">2</span> atm) and temperatures (973-1173 K) to investigate the condition of Si crystal formation from <span class="hlt">Na</span>Si by <span class="hlt">Na</span> evaporation. Silicon single crystals 1-3 mm in diameter were grown by evaporation of <span class="hlt">Na</span> from <span class="hlt">Na</span>-Si melt at 1173 K and p<span class="hlt">Na</span>=0.74 atm.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2001JMoSt.596..171R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2001JMoSt.596..171R"><span>The influence of <span class="hlt">Na</span> + and Ca <span class="hlt">2</span>+ ions on the SiO <span class="hlt">2</span>-AlPO 4 materials structure — IR and Raman studies</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Rokita, M.; Mozgawa, W.; Handke, M.</p> <p>2001-09-01</p> <p>The series of samples containing 0-20 mol% of <span class="hlt">Na</span>CaPO4 and 20-0 mol% of AlPO4, respectively, with the constant amount of SiO<span class="hlt">2</span> (80 mol%) have been selected. The materials were prepared using both sol-gel as well as aerosil pseudo-aqua solution method. The AlPO4·SiO<span class="hlt">2</span> and <span class="hlt">Na</span>CaPO4·SiO<span class="hlt">2</span> (80 mol% of SiO<span class="hlt">2</span>) samples have been prepared. IR and Raman spectra of these samples are presented. The spectra of materials from <span class="hlt">Na</span>CaPO4-AlPO4-SiO<span class="hlt">2</span> system are compared to those of <span class="hlt">Na</span>CaPO4·SiO<span class="hlt">2</span> and AlPO4·SiO<span class="hlt">2</span> sample (samples without Al3+ or <span class="hlt">Na</span>+ and Ca<span class="hlt">2</span>+ cations, respectively). The studies have enabled us to identify the bands arising from the internal and lattice vibrations. The slight differences between the spectra of sol-gel and aerosil pseudo-aqua solution materials are pointed out and discussed. The influence of <span class="hlt">Na</span>+ and Ca<span class="hlt">2</span>+ ions on the AlPO4-SiO<span class="hlt">2</span> materials structure is analysed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4971830','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4971830"><span><span class="hlt">Na</span>+-independent phosphate transport in Caco<span class="hlt">2</span>BBE 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>Candeal, Eduardo; Caldas, Yupanqui A.; Guillén, Natalia; Levi, Moshe</p> <p>2014-01-01</p> <p>Pi transport in epithelia has both <span class="hlt">Na</span>+-dependent and <span class="hlt">Na</span>+-independent components, but so far only <span class="hlt">Na</span>+-dependent transporters have been characterized in detail and molecularly identified. Consequently, in the present study, we initiated the characterization and analysis of intestinal <span class="hlt">Na</span>+-independent Pi transport using an in vitro model, Caco<span class="hlt">2</span>BBE cells. Only <span class="hlt">Na</span>+-independent Pi uptake was observed in these cells, and Pi uptake was dramatically increased when cells were incubated in high-Pi DMEM (4 mM) from 1 day to several days. No response to low-Pi medium was observed. The increased Pi transport was mainly caused by Vmax changes, and it was prevented by actinomycin D and cycloheximide. Pi transport in cells grown in 1 mM Pi (basal DMEM) decreased at pH > 7.5, and it was inhibited with proton ionophores. Pi transport in cells incubated with 4 mM Pi increased with alkaline pH, suggesting a preference for divalent phosphate. Pi uptake in cells in 1 mM Pi was completely inhibited only by Pi and partially inhibited by phosphonoformate, oxalate, DIDS, SITS, SO42−, HCO3−, and arsenate. This inhibition pattern suggests that more than one Pi transporter is active in cells maintained with 1 mM Pi. Phosphate transport from cells maintained at 4 mM Pi was only partially inhibited by phosphonoformate, oxalate, and arsenate. Attempts to identify the responsible transporters showed that multifunctional anion exchangers of the Slc26 family as well as members of Slc17, Slc20, and Slc37 and the Pi exporter xenotropic and polytropic retrovirus receptor 1 are not involved. PMID:25298422</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018AIPC.1953e0007C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AIPC.1953e0007C"><span>X-ray diffraction and infrared spectroscopy studies of Ba(Fe1/<span class="hlt">2</span>Nb1/<span class="hlt">2</span>)O3-(<span class="hlt">Na</span>1/<span class="hlt">2</span>Bi1/<span class="hlt">2</span>)TiO3 ceramics</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Chandra, K. P.; Yadav, Anjana; Prasad, K.</p> <p>2018-05-01</p> <p>Ceramics (1-x)Ba(Fe1/<span class="hlt">2</span>Nb1/<span class="hlt">2</span>)O3-x(<span class="hlt">Na</span>1/<span class="hlt">2</span>Bi1/<span class="hlt">2</span>)TiO3; 0≤x≤1.0 were prepared by conventional ceramic synthesis technique. Rietveld refinements of X-ray diffraction data of these ceramics were carried out using FullProf software and determined their crystal symmetry, space group and unit cell dimensions. Rietveld refinement revealed that Ba(Fe1/<span class="hlt">2</span>Nb1/<span class="hlt">2</span>)O3 has cubic structure with space group Pm 3 ¯ m and <span class="hlt">Na</span>1/<span class="hlt">2</span>Bi1/<span class="hlt">2</span>)TiO3 has rhombohedral structure with space group R3c. Addition of (<span class="hlt">Na</span>1/<span class="hlt">2</span>Bi1/<span class="hlt">2</span>)TiO3 to Ba(Fe1/<span class="hlt">2</span>Nb1/<span class="hlt">2</span>)O3 resulted in the change of unit cell structure from cubic to tetragonal (P4/mmm) for x = 0.75 and the X-Ray diffraction peaks slightly shift towards higher Bragg's angle, suggesting slight decrease in unit cell volume. SEM studies were carried out in order to access the quality of the prepared ceramics which showed a change in grain shapes with the increase of (<span class="hlt">Na</span>1/<span class="hlt">2</span>Bi1/<span class="hlt">2</span>)TiO3 content. FTIR spectra confirmed the formation of perovskite type solid solutions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1185906','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1185906"><span>Divalent europium doped and un-doped calcium iodide scintillators: Scintillator characterization and single crystal growth</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>Boatner, L. A.; Ramey, J. O.; Kolopus, J. A.</p> <p>2015-02-21</p> <p>Initially, the alkaline-earth scintillator, <span class="hlt">CaI</span> <span class="hlt">2</span>:Eu <span class="hlt">2</span>+, was discovered around 1964 by Hofstadter, Odell, and Schmidt. Serious practical problems quickly arose, however, that were associated with the growth of large monolithic single crystals of this material due to its lamellar, mica-like structure. As a result of its theoretically higher light yield, <span class="hlt">CaI</span> <span class="hlt">2</span>:Eu <span class="hlt">2</span>+ has the potential to exceed the excellent scintillation performance of SrI <span class="hlt">2</span>:Eu <span class="hlt">2</span>+. In fact, theoretical predictions for the light yield of <span class="hlt">CaI</span><span class="hlt">2</span>:Eu <span class="hlt">2</span>+ scintillators suggested that an energy resolution approaching <span class="hlt">2</span>% at 662 keV could be achievable. Like the early SrI <span class="hlt">2</span>:Eu <span class="hlt">2</span>+ scintillator, themore » performance of <span class="hlt">CaI</span> <span class="hlt">2</span>:Eu <span class="hlt">2</span>+ scintillators has traditionally suffered due, at least in part, to outdated materials synthesis, component stoichiometry/purity, and single-crystal-growth techniques. Based on our recent work on SrI <span class="hlt">2</span>:Eu <span class="hlt">2</span>+ scintillators in single-crystal form, we have developed new techniques that are applied here to <span class="hlt">CaI</span> <span class="hlt">2</span>:Eu <span class="hlt">2</span>+ and pure <span class="hlt">CaI</span> <span class="hlt">2</span> with the goal of growing large un-cracked crystals and, potentially, realizing the theoretically predicted performance of the <span class="hlt">CaI</span> <span class="hlt">2</span>:Eu <span class="hlt">2</span>+ form of this material. Calcium iodide does not adhere to modern glassy carbon Bridgman crucibles - so there should be no differential thermal-contraction-induced crystal/crucible stresses on cooling that would result in crystal cracking of the lamellar structure of <span class="hlt">CaI</span> <span class="hlt">2</span>. Here we apply glassy carbon crucible Bridgman growth, high-purity growth-charge compounds, our molten salt processing/filtration technique, and extended vacuum-melt-pumping methods to the growth of both <span class="hlt">CaI</span> <span class="hlt">2</span>:Eu <span class="hlt">2</span>+ and un-doped <span class="hlt">CaI</span> <span class="hlt">2</span>. Moreover, large scintillating single crystals were obtained, and detailed characterization studies of the scintillation properties of <span class="hlt">CaI</span> <span class="hlt">2</span>:Eu <span class="hlt">2</span>+ and pure <span class="hlt">CaI</span> <span class="hlt">2</span> single crystals are presented that include studies of the effects of plastic deformation of the crystals on the scintillator performance.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22462133-sg2na-enhances-cancer-cell-survival-stabilizing-dj-andthusactivating-akt','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22462133-sg2na-enhances-cancer-cell-survival-stabilizing-dj-andthusactivating-akt"><span>SG<span class="hlt">2</span><span class="hlt">NA</span> enhances cancer cell survival by stabilizing DJ-1 and thus activating Akt</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>Tanti, Goutam Kumar, E-mail: goutamjnu@hotmail.com; Pandey, Shweta; Goswami, Shyamal K.</p> <p>2015-08-07</p> <p>SG<span class="hlt">2</span><span class="hlt">NA</span> in association with striatin and zinedin forms a striatin family of WD-40 repeat proteins. This family of proteins functions as scaffold in different signal transduction pathways. They also act as a regulatory subunit of protein phosphatase <span class="hlt">2</span>A. We have shown that SG<span class="hlt">2</span><span class="hlt">NA</span> which evolved first in the metazoan evolution among the striatin family members expresses different isoforms generated out of alternative splicing. We have also shown that SG<span class="hlt">2</span><span class="hlt">NA</span> protects cells from oxidative stress by recruiting DJ-1 and Akt to mitochondria and membrane in the post-mitotic neuronal cells. DJ-1 is both cancer and Parkinson's disease related protein. In the presentmore » study we have shown that SG<span class="hlt">2</span><span class="hlt">NA</span> protects DJ-1 from proteasomal degradation in cancer cells. Hence, downregulation of SG<span class="hlt">2</span><span class="hlt">NA</span> reduces DJ-1/Akt colocalization in cancer cells resulting in the reduction of anchorage dependent and independent growth. Thus SG<span class="hlt">2</span><span class="hlt">NA</span> enhances cancer cell survival. Reactive oxygen species enhances SG<span class="hlt">2</span><span class="hlt">NA</span>, DJ-1 and Akt trimerization. Removal of the reactive oxygen species by N-acetyl-cysteine thus reduces cancer cell growth. - Highlights: • Reactive oxygen species (ROS) play potential role in cancer cell proliferation. • It enhances the association between DJ-1 and Akt mediated by SG<span class="hlt">2</span><span class="hlt">NA</span>. • In cancer cells SG<span class="hlt">2</span><span class="hlt">NA</span> stabilizes DJ-1 by inhibiting it from proteosomal degradation. • DJ-1 then activates Akt and cancer cells get their property of enhanced proliferation by sustained activation of Akt. • Further study on this field could lead to new target for cancer therapy.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26209057','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26209057"><span>Arterial α<span class="hlt">2</span>-<span class="hlt">Na</span>+ pump expression influences blood pressure: lessons from novel, genetically engineered smooth muscle-specific α<span class="hlt">2</span> mice.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Chen, Ling; Song, Hong; Wang, Youhua; Lee, Jane C; Kotlikoff, Michael I; Pritchard, Tracy J; Paul, Richard J; Zhang, Jin; Blaustein, Mordecai P</p> <p>2015-09-01</p> <p>Arterial myocytes express α1-catalytic subunit isoform <span class="hlt">Na</span>(+) pumps (75-80% of total), which are ouabain resistant in rodents, and high ouabain affinity α<span class="hlt">2</span>-<span class="hlt">Na</span>(+) pumps. Mice with globally reduced α<span class="hlt">2</span>-pumps (but not α1-pumps), mice with mutant ouabain-resistant α<span class="hlt">2</span>-pumps, and mice with a smooth muscle (SM)-specific α<span class="hlt">2</span>-transgene (α<span class="hlt">2</span> (SM-Tg)) that induces overexpression all have altered blood pressure (BP) phenotypes. We generated α<span class="hlt">2</span> (SM-DN) mice with SM-specific α<span class="hlt">2</span> (not α1) reduction (>50%) using nonfunctional dominant negative (DN) α<span class="hlt">2</span>. We compared α<span class="hlt">2</span> (SM-DN) and α<span class="hlt">2</span> (SM-Tg) mice to controls to determine how arterial SM α<span class="hlt">2</span>-pumps affect vasoconstriction and BP. α<span class="hlt">2</span> (SM-DN) mice had elevated basal mean BP (mean BP by telemetry: 117 ± 4 vs. 106 ± 1 mmHg, n = 7/7, P < 0.01) and enhanced BP responses to chronic ANG II infusion (240 ng·kg(-1)·min(-1)) and high (6%) <span class="hlt">Na</span>Cl. Several arterial Ca(<span class="hlt">2</span>+) transporters, including <span class="hlt">Na</span>(+)/Ca(<span class="hlt">2</span>+) exchanger 1 (NCX1) and sarcoplasmic reticulum and plasma membrane Ca(<span class="hlt">2</span>+) pumps [sarco(endo)plasmic reticulum Ca(<span class="hlt">2</span>+)-ATPase <span class="hlt">2</span> (SERCA<span class="hlt">2</span>) and plasma membrane Ca(<span class="hlt">2</span>+)-ATPase 1 (PMCA1)], were also reduced (>50%). α<span class="hlt">2</span> (SM-DN) mouse isolated small arteries had reduced myogenic reactivity, perhaps because of reduced Ca(<span class="hlt">2</span>+) transporter expression. In contrast, α<span class="hlt">2</span> (SM-Tg) mouse aortas overexpressed α<span class="hlt">2</span> (><span class="hlt">2</span>-fold), NCX1, SERCA<span class="hlt">2</span>, and PMCA1 (43). α<span class="hlt">2</span> (SM-Tg) mice had reduced basal mean BP (104 ± 1 vs. 109 ± <span class="hlt">2</span> mmHg, n = 15/9, P < 0.02) and attenuated BP responses to chronic ANG II (300-400 ng·kg(-1)·min(-1)) with or without <span class="hlt">2</span>% <span class="hlt">Na</span>Cl but normal myogenic reactivity. NCX1 expression was inversely related to basal BP in SM-α<span class="hlt">2</span> engineered mice but was directly related in SM-NCX1 engineered mice. NCX1, which usually mediates arterial Ca(<span class="hlt">2</span>+) entry, and α<span class="hlt">2</span>-<span class="hlt">Na</span>(+) pumps colocalize at plasma membrane-sarcoplasmic reticulum junctions and functionally couple via the local <span class="hlt">Na</span>(+) gradient to help regulate cell Ca(<span class="hlt">2</span>+). Altered Ca(<span class="hlt">2</span>+) transporter expression in</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29554219','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29554219"><span>PRRT<span class="hlt">2</span> controls neuronal excitability by negatively modulating <span class="hlt">Na</span>+ channel 1.<span class="hlt">2</span>/1.6 activity.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Fruscione, Floriana; Valente, Pierluigi; Sterlini, Bruno; Romei, Alessandra; Baldassari, Simona; Fadda, Manuela; Prestigio, Cosimo; Giansante, Giorgia; Sartorelli, Jacopo; Rossi, Pia; Rubio, Alicia; Gambardella, Antonio; Nieus, Thierry; Broccoli, Vania; Fassio, Anna; Baldelli, Pietro; Corradi, Anna; Zara, Federico; Benfenati, Fabio</p> <p>2018-04-01</p> <p>See Lerche (doi:10.1093/brain/awy073) for a scientific commentary on this article.Proline-rich transmembrane protein <span class="hlt">2</span> (PRRT<span class="hlt">2</span>) is the causative gene for a heterogeneous group of familial paroxysmal neurological disorders that include seizures with onset in the first year of life (benign familial infantile seizures), paroxysmal kinesigenic dyskinesia or a combination of both. Most of the PRRT<span class="hlt">2</span> mutations are loss-of-function leading to haploinsufficiency and 80% of the patients carry the same frameshift mutation (c.649dupC; p.Arg217Profs*8), which leads to a premature stop codon. To model the disease and dissect the physiological role of PRRT<span class="hlt">2</span>, we studied the phenotype of neurons differentiated from induced pluripotent stem cells from previously described heterozygous and homozygous siblings carrying the c.649dupC mutation. Single-cell patch-clamp experiments on induced pluripotent stem cell-derived neurons from homozygous patients showed increased <span class="hlt">Na</span>+ currents that were fully rescued by expression of wild-type PRRT<span class="hlt">2</span>. Closely similar electrophysiological features were observed in primary neurons obtained from the recently characterized PRRT<span class="hlt">2</span> knockout mouse. This phenotype was associated with an increased length of the axon initial segment and with markedly augmented spontaneous and evoked firing and bursting activities evaluated, at the network level, by multi-electrode array electrophysiology. Using HEK-293 cells stably expressing Nav channel subtypes, we demonstrated that the expression of PRRT<span class="hlt">2</span> decreases the membrane exposure and <span class="hlt">Na</span>+ current of Nav1.<span class="hlt">2</span>/Nav1.6, but not Nav1.1, channels. Moreover, PRRT<span class="hlt">2</span> directly interacted with Nav1.<span class="hlt">2</span>/Nav1.6 channels and induced a negative shift in the voltage-dependence of inactivation and a slow-down in the recovery from inactivation. In addition, by co-immunoprecipitation assays, we showed that the PRRT<span class="hlt">2</span>-Nav interaction also occurs in brain tissue. The study demonstrates that the lack of PRRT<span class="hlt">2</span> leads to a hyperactivity of voltage</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5888929','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5888929"><span>PRRT<span class="hlt">2</span> controls neuronal excitability by negatively modulating <span class="hlt">Na</span>+ channel 1.<span class="hlt">2</span>/1.6 activity</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Fruscione, Floriana; Valente, Pierluigi; Sterlini, Bruno; Romei, Alessandra; Baldassari, Simona; Fadda, Manuela; Prestigio, Cosimo; Giansante, Giorgia; Sartorelli, Jacopo; Rossi, Pia; Rubio, Alicia; Gambardella, Antonio; Nieus, Thierry; Broccoli, Vania; Fassio, Anna; Baldelli, Pietro; Corradi, Anna; Zara, Federico</p> <p>2018-01-01</p> <p>Abstract See Lerche (doi:10.1093/brain/awy073) for a scientific commentary on this article. Proline-rich transmembrane protein <span class="hlt">2</span> (PRRT<span class="hlt">2</span>) is the causative gene for a heterogeneous group of familial paroxysmal neurological disorders that include seizures with onset in the first year of life (benign familial infantile seizures), paroxysmal kinesigenic dyskinesia or a combination of both. Most of the PRRT<span class="hlt">2</span> mutations are loss-of-function leading to haploinsufficiency and 80% of the patients carry the same frameshift mutation (c.649dupC; p.Arg217Profs*8), which leads to a premature stop codon. To model the disease and dissect the physiological role of PRRT<span class="hlt">2</span>, we studied the phenotype of neurons differentiated from induced pluripotent stem cells from previously described heterozygous and homozygous siblings carrying the c.649dupC mutation. Single-cell patch-clamp experiments on induced pluripotent stem cell-derived neurons from homozygous patients showed increased <span class="hlt">Na</span>+ currents that were fully rescued by expression of wild-type PRRT<span class="hlt">2</span>. Closely similar electrophysiological features were observed in primary neurons obtained from the recently characterized PRRT<span class="hlt">2</span> knockout mouse. This phenotype was associated with an increased length of the axon initial segment and with markedly augmented spontaneous and evoked firing and bursting activities evaluated, at the network level, by multi-electrode array electrophysiology. Using HEK-293 cells stably expressing Nav channel subtypes, we demonstrated that the expression of PRRT<span class="hlt">2</span> decreases the membrane exposure and <span class="hlt">Na</span>+ current of Nav1.<span class="hlt">2</span>/Nav1.6, but not Nav1.1, channels. Moreover, PRRT<span class="hlt">2</span> directly interacted with Nav1.<span class="hlt">2</span>/Nav1.6 channels and induced a negative shift in the voltage-dependence of inactivation and a slow-down in the recovery from inactivation. In addition, by co-immunoprecipitation assays, we showed that the PRRT<span class="hlt">2</span>-Nav interaction also occurs in brain tissue. The study demonstrates that the lack of PRRT<span class="hlt">2</span> leads to a hyperactivity of</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26618389','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26618389"><span>Composition Dependence of the <span class="hlt">Na</span>(+) Ion Conductivity in 0.5<span class="hlt">Na</span><span class="hlt">2</span>S + 0.5[xGeS<span class="hlt">2</span> + (1 - x)PS5/<span class="hlt">2</span>] Mixed Glass Former Glasses: A Structural Interpretation of a Negative Mixed Glass Former Effect.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Martin, Steve W; Bischoff, Christian; Schuller, Katherine</p> <p>2015-12-24</p> <p>A negative mixed glass former effect (MGFE) in the <span class="hlt">Na</span>(+) ion conductivity of glass has been found in 0.5<span class="hlt">Na</span><span class="hlt">2</span>S + 0.5[xGeS<span class="hlt">2</span> + (1 - x)PS5/<span class="hlt">2</span>] glasses where the <span class="hlt">Na</span>(+) ion conductivity is significantly smaller for all of the ternary glasses than either of the binary end-member glasses. The minimum conductivity of ∼0.4 × 10(-6) (Ω cm)(-1) at 25 °C occurs for the x = 0.7 glass. Prior to this observation, the alkali ion conductivity of sulfide glasses at constant alkali concentration, but variable ratio of one glass former for another (x) ternary mixed glass former (MGF) glasses, has always produced a positive MGFE in the alkali ion conductivity; that is, the ternary glasses have always had higher ion conductivities that either of the end-member binary glasses. While the <span class="hlt">Na</span>(+) ion conductivity exhibits a single global minimum value, the conductivity activation energy exhibits a bimodal double maximum at x ≈ 0.4 and x ≈ 0.7. The modified Christensen-Martin-Anderson-Stuart (CMAS) model of the activation energies reveals the origin of the negative MGFE to be due to an increase in the dielectric stiffness (a decrease in relative dielectric permittivity) of these glasses. When coupled with an increase in the average <span class="hlt">Na</span>(+) ion jump distance and a slight increase in the mechanical stiffness of the glass, this causes the activation energy to go through maximum values and thereby produce the negative MGFE. The double maximum in the conductivity activation energy is coincident with double maximums in CMAS calculated strain, ΔES, and Coulombic, ΔEC, activation energies. In these ternary glasses, the increase in the dielectric stiffness of the glass arises from a negative deviation of the limiting high frequency dielectric permittivity as compared to the binary end-member glasses. While the CMAS calculated total activation energies ΔEact = ΔES + ΔEC are found to reproduce the overall shape of the composition dependence of the measured ΔEact values, they are consistently</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018CPL...701..147L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018CPL...701..147L"><span>Bulk and grain-boundary ionic conductivity in sodium zirconophosphosilicate <span class="hlt">Na</span>3Zr<span class="hlt">2</span>(SiO4)<span class="hlt">2</span>PO4 (NASICON)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lunghammer, S.; Ma, Q.; Rettenwander, D.; Hanzu, I.; Tietz, F.; Wilkening, H. M. R.</p> <p>2018-06-01</p> <p>Sodium zirconophosphosilicates (<span class="hlt">Na</span>1+x Zr<span class="hlt">2</span>(P1-x SixO4)3 (0 < x < 3)) currently experience a kind of renaissance as promising ceramic electrolytes for safe all-solid-state <span class="hlt">Na</span> batteries. Such energy storage systems are an emerging option for next-generation technologies with attractive cost due to the use of abundant elements as sodium. To identify the right candidates their ion transport properties need to be precisely studied. In many cases less is known about the contributions of blocking grain boundaries to the overall charge carrier transport. Here, we took advantage of broadband impedance and conductivity spectroscopy carried out at sufficiently low temperature to make visible these two contributions for polycrystalline <span class="hlt">Na</span>3Zr<span class="hlt">2</span>(SiO4)<span class="hlt">2</span>PO4. It turned out that ion transport across the grain boundaries of a sintered pellet do not greatly hinder long-range ion dynamics. While bulk ion dynamics in <span class="hlt">Na</span>3Zr<span class="hlt">2</span>(SiO4)<span class="hlt">2</span>PO4 is characterized by 1.0 mS cm-1, the grain boundary ionic conductivity is only slightly lower viz. 0.7 mS cm-1. The latter value is of large practical interest as it allows the realization of all-solid-state <span class="hlt">Na</span> batteries without strong interfering resistances from grain boundaries.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2009PhPro...2.1433N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009PhPro...2.1433N"><span>Elaboration and characterization of solid materials of types zeolite <span class="hlt">Na</span>A and faujasite <span class="hlt">Na</span>Y exchanged by zinc metallic ions Zn<span class="hlt">2</span>+</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Nibou, D.; Amokrane, S.; Mekatel, H.; Lebaili, N.</p> <p>2009-11-01</p> <p>The present work deals with the elaborated of <span class="hlt">Na</span>A and faujasite <span class="hlt">Na</span>Y solid materials according to a hydrothermal crystallization of amorphous gels composed of solutions of silicon, aluminum and sodium. The process elaboration has been achieved in autoclaves made of steel lined in Teflon under different operating conditions of temperature of heating, time of contact and stirring. After crystallization, the samples were characterized by different techniques such as X ray diffraction, scanning electronic microscopy, infrared spectroscopy, thermal analysis, and chemical analysis. Pure solid materials <span class="hlt">Na</span>A and <span class="hlt">Na</span>Y zeolites were obtained and were impregnated by (Zn<span class="hlt">2</span>+) ions by ion exchange process. The effects of various parameters such as initial metal concentration, pH, solid-liquid ratio (R) and temperature on the exchange percentage are studied. The equilibrium isotherms of zinc ions sorption are also evaluated using Langmuir and Freundlich models. Thermodynamic parameters, i.e. enthalpy of adsorption ΔHads∘, entropy change ΔSads∘ and Gibbs free energy ΔGads∘ for the sorption of zinc ions on <span class="hlt">Na</span>A and <span class="hlt">Na</span>Y zeolites were examined.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018CPL...700..122Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018CPL...700..122Y"><span>A new global analytical potential energy surface of <span class="hlt">Na</span>H<span class="hlt">2</span>+ system and dynamical calculation for H(<span class="hlt">2</span>S) + <span class="hlt">Na</span>H+(X<span class="hlt">2</span>Σ+) → <span class="hlt">Na</span>+(1S) + H<span class="hlt">2</span>(X1Σg+) reaction</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yuan, Meiling; Li, Wentao; Yuan, Jiuchuang</p> <p>2018-05-01</p> <p>A new global potential energy surface (PES) of the <span class="hlt">Na</span>H<span class="hlt">2</span>+ system is constructed by fitting 27,621 ab initio energy points with the neural network method. The root mean square error of the new PES is only 4.1609 × 10-4 eV. Based on the new PES, dynamical calculations have been performed using the time-dependent quantum wave packet method. These results are then compared with the H(<span class="hlt">2</span>S) + LiH+(X<span class="hlt">2</span>Σ+) → Li+(1S) + H<span class="hlt">2</span>(X1Σg+) reaction. The direct abstract mechanism is found to play an important role in the reaction because only forward scattering signals on the differential cross section results for all calculated collision energies.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28791103','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28791103"><span>Asymmetric supercapacitors with high energy density based on helical hierarchical porous <span class="hlt">Na</span> x MnO<span class="hlt">2</span> and MoO<span class="hlt">2</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Lu, Xue-Feng; Huang, Zhi-Xiang; Tong, Ye-Xiang; Li, Gao-Ren</p> <p>2016-01-01</p> <p>Helical hierarchical porous <span class="hlt">Na</span> x MnO <span class="hlt">2</span> /CC and MoO <span class="hlt">2</span> /CC, which are assembled from nanosheets and nanoparticles, respectively, are fabricated using a simple electrodeposition method. These unique helical porous structures enable electrodes to have a high capacitance and an outstanding cycling performance. Based on the helical <span class="hlt">Na</span> x MnO <span class="hlt">2</span> /CC as the positive electrodes and helical MoO <span class="hlt">2</span> /CC as the negative electrodes, high performance <span class="hlt">Na</span> x MnO <span class="hlt">2</span> /CC//MoO <span class="hlt">2</span> /CC asymmetric supercapacitors (ASCs) are successfully assembled, and they achieve a maximum volume C sp of <span class="hlt">2</span>.04 F cm -3 and a maximum energy density of 0.92 mW h cm -3 for the whole device and an excellent cycling stability with 97.22% C sp retention after 6000 cycles.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2007JSSCh.180.1934K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2007JSSCh.180.1934K"><span>Structural studies on a high-pressure polymorph of <span class="hlt">Na</span>YSi <span class="hlt">2</span>O 6</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kahlenberg, Volker; Konzett, Jürgen; Kaindl, Reinhard</p> <p>2007-06-01</p> <p>High-pressure synthesis experiments in the system <span class="hlt">Na</span> <span class="hlt">2</span>O-Y <span class="hlt">2</span>O 3-SiO <span class="hlt">2</span> revealed the existence of a previously unknown polymorph of <span class="hlt">Na</span>YSi <span class="hlt">2</span>O 6 or <span class="hlt">Na</span> 3Y 3[Si 3O 9] <span class="hlt">2</span> which was quenched from 3.0 GPa and 1000 °C. Structural investigations on this modification have been performed using single-crystal X-ray diffraction data collected at ambient conditions. Furthermore, unpolarized micro-Raman spectra have been obtained from single-crystal material. The high-P modification of <span class="hlt">Na</span>YSi <span class="hlt">2</span>O 6 crystallizes in the centrosymmetric space group C<span class="hlt">2</span>/ c with 12 formula units per cell ( a=8.2131(9) Å, b=10.3983(14) Å, c=17.6542(21) Å, β=100.804(9)°, V=1481.0(3) Å 3, R(| F|)=0.033 for 1142 independent observed reflections) and belongs to the group of cyclo-silicates. Basic building units are isolated three-membered [Si 3O 9] rings located in layers parallel to (010). Within a single layer the rings are concentrated in strings parallel to [100]. The sequence of directedness of up ( U) or down ( D) pointing tetrahedra of a single ring is UUU or DDD, respectively. Stacking of the layers parallel to b results in the formation of a three-dimensional structure in which yttrium and sodium cations are incorporated for charge compensation. In more detail, four non-tetrahedral cation positions can be differentiated which are coordinated by 6 and 8 oxygen ligands. Refinements of the site occupancies did not reveal any indication for mixed <span class="hlt">Na</span>-Y populations on these positions. Finally, several geometrical parameters of rings occurring in cyclo-trisilicate structures have been compiled and are discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26877029','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26877029"><span>Preparation and photoelectrocatalytic performance of N-doped TiO<span class="hlt">2</span>/<span class="hlt">Na</span>Y zeolite membrane composite electrode material.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Cheng, Zhi-Lin; Han, Shuai</p> <p>2016-01-01</p> <p>A novel composite electrode material based on a N-doped TiO<span class="hlt">2</span>-loaded <span class="hlt">Na</span>Y zeolite membrane (N-doped TiO<span class="hlt">2</span>/<span class="hlt">Na</span>Y zeolite membrane) for photoelectrocatalysis was presented. X-ray diffraction (XRD), scanning electron microscopy (SEM), UV-visible (UV-vis) and X-ray photoelectron spectroscopy (XPS) characterization techniques were used to analyze the structure of the N-doped TiO<span class="hlt">2</span>/<span class="hlt">Na</span>Y zeolite membrane. The XRD and SEM results verified that the N-doped TiO<span class="hlt">2</span> nanoparticles with the size of ca. 20 nm have been successfully loaded on the porous stainless steel-supported <span class="hlt">Na</span>Y zeolite membrane. The UV-vis result showed that the N-doped TiO<span class="hlt">2</span>/<span class="hlt">Na</span>Y zeolite membrane exhibited a more obvious red-shift than that of N-TiO<span class="hlt">2</span> nanoparticles. The XPS characterization revealed that the doping of N element into TiO<span class="hlt">2</span> was successfully achieved. The photoelectrocatalysis performance of the N-doped TiO<span class="hlt">2</span>/<span class="hlt">Na</span>Y zeolite membrane composite electrode material was evaluated by phenol removal and also the effects of reaction conditions on the catalytic performance were investigated. Owing to exhibiting an excellent catalytic activity and good recycling stability, the N-doped TiO<span class="hlt">2</span>/<span class="hlt">Na</span>Y zeolite membrane composite electrode material was of promising application for photoelectrocatalysis in wastewater treatment.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1226014','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1226014"><span>Understanding the Effect of <span class="hlt">Na</span> in Improving the Performance of CuInSe <span class="hlt">2</span> Based Photovoltaics</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>Dobson, Kevin D.</p> <p></p> <p>Cu(In,Ga)Se <span class="hlt">2</span> (CIGS) thin film photovoltaic technology is in the early stages of commercialization with an annual manufacturing capacity over 1 GW and has demonstrated the highest module efficiency of any of the thin film technologies. However there still is a lack of fundamental understanding of the relationship between the material properties and solar cell device operation. It is well known that the incorporation of a small amount of <span class="hlt">Na</span> into the CIGS film during processing is essential for high efficiency devices. However, there are conflicting explanations for how <span class="hlt">Na</span> behaves at the atomic scale. This report investigates how Namore » is incorporated into the CIGS device structure and evaluates the diffusion of <span class="hlt">Na</span> into CIGS grain boundaries (GBs) and bulk crystallites. Participants: This project was carried out at the Institute of Energy Conversion at the University of Delaware, collaborating with the Rockett group at the University of Illinois Urbana-Champagne. Significant Findings: The significant outcomes of this project for each task include; Task 1.0: Effect of <span class="hlt">Na</span> in Devices Fabricated on PVD Deposited CIGS; <span class="hlt">Na</span> diffusion occurs through the Mo back contact via GBs driven by the presence of oxygen; <span class="hlt">Na</span> reversibly compensates donor defects in CIGS GBs,Task <span class="hlt">2</span>.0: <span class="hlt">Na</span> Incorporation in Single Crystal CIGS; and bulk <span class="hlt">Na</span> diffusion proceeds rapidly such that grains are <span class="hlt">Na</span>-saturated immediately following CIGS thin film manufacture. Industry Guidance: The presented results offer interesting concepts for modification of manufacturing processes of CIGS-based PV modules. Possible approaches to improve control of <span class="hlt">Na</span> uptake and uniformly increase levels in CIGS films are highlighted for processes that employ either soda-lime glass or <span class="hlt">Na</span>F as the <span class="hlt">Na</span> source. Concepts include the potential of O <span class="hlt">2</span> or oxidative based treatments of Mo back contacts to improve <span class="hlt">Na</span> diffusion through the metal film and increase <span class="hlt">Na</span> uptake into the growing CIGS. This project has also offered fundamental</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li class="active"><span>21</span></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_21 --> <div id="page_22" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li class="active"><span>22</span></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li><a href="#" onclick='return showDiv("page_24");'>24</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="421"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://eric.ed.gov/?q=POLYMER&pg=6&id=EJ921265','ERIC'); return false;" href="https://eric.ed.gov/?q=POLYMER&pg=6&id=EJ921265"><span>"JCE" Classroom Activity Connections: <span class="hlt">Na</span>Cl or CaCl[subscript <span class="hlt">2</span>], Smart Polymer Gel Tells More</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>Chen, Yueh-Huey; Lin, Jia-Ying; Wang, Yu-Chen; Yaung, Jing-Fun</p> <p>2010-01-01</p> <p>This classroom activity connection demonstrates the differences between the effects of <span class="hlt">Na</span>Cl (a salt of monovalent metal ions) and CaCl[subscript <span class="hlt">2</span>] (a salt of polyvalent metal ions) on swollen superabsorbent polymer gels. Being ionic compounds, <span class="hlt">Na</span>Cl and CaCl[subscript <span class="hlt">2</span>] both collapse the swollen polymer gels. The gel contracted by <span class="hlt">Na</span>Cl reswells…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/973864','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/973864"><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>Matzel, J P; Jacobsen, B; Hutcheon, I D</p> <p></p> <p>The abundance of short-lived radionuclides (SLRs) in early solar system materials provide key information about their nucleosynthetic origin and can constrain the timing of early solar system events. Excesses of {sup 36}S ({sup 36}S*) correlated with {sup 35}Cl/{sup 34}S ratios provide direct evidence for in situ decay of {sup 36}Cl ({tau}{sub 1/<span class="hlt">2</span>} {approx} 0.3 Ma) and have been reported in sodalite (<span class="hlt">Na</span>{sub 8}Al{sub 6}Si{sub 6}O{sub 24}Cl{sub <span class="hlt">2</span>}) and wadalite (Ca{sub 6}Al{sub 5}Si{sub <span class="hlt">2</span>}O{sub 16}Cl{sub 3}) in <span class="hlt">CAIs</span> and chondrules from the Allende and Ningqiang CV carbonaceous chondrites. While previous studies demonstrate unequivocally that {sup 36}Cl was extant in the earlymore » solar system, no consensus on the origin or initial abundance of {sup 36}Cl has emerged. Understanding the origin of {sup 36}Cl, as well as the reported variation in the initial {sup 36}Cl/{sup 35}Cl ratio, requires addressing when, where and how chlorine was incorporated into <span class="hlt">CAIs</span> and chondrules. These factors are key to distinguishing between stellar nucleosynthesis or energetic particle irradiation for the origin of {sup 36}Cl. Wadalite is a chlorine-rich secondary mineral with structural and chemical affinities to grossular. The high chlorine ({approx}12 wt%) and very low sulfur content (<<0.01 wt%) make wadalite ideal for studies of the {sup 36}Cl-{sup 36}S system. Wadalite is present in Allende <span class="hlt">CAIs</span> exclusively in the interior regions either in veins crosscutting melilite or in zones between melilite and anorthite associated with intergrowths of grossular, monticellite, and wollastonite. Wadalite and sodalite most likely resulted from open-system alteration of primary minerals with a chlorine-rich fluid phase. We recently reported large {sup 36}S* correlated with {sup 35}Cl/{sup 34}S in wadalite in Allende Type B <span class="hlt">CAI</span> AJEF, yielding a ({sup 36}Cl/{sup 35}Cl){sub 0} ratio of (1.7 {+-} 0.3) x 10{sup -5}. This value is the highest reported {sup 36}Cl/{sup 35}Cl ratio and is {approx}5</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/19792680','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/19792680"><span>One-dimensional magnetic fluctuations in the spin-<span class="hlt">2</span> triangular lattice alpha-<span class="hlt">Na</span>MnO<span class="hlt">2</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Stock, C; Chapon, L C; Adamopoulos, O; Lappas, A; Giot, M; Taylor, J W; Green, M A; Brown, C M; Radaelli, P G</p> <p>2009-08-14</p> <p>The S=<span class="hlt">2</span> anisotropic triangular lattice alpha-<span class="hlt">Na</span>MnO<span class="hlt">2</span> is studied by neutron inelastic scattering. Antiferromagnetic order occurs at T< or =45 K with opening of a spin gap. The spectral weight of the magnetic dynamics above the gap (Delta approximately equal to 7.5 meV) has been analyzed by the single-mode approximation. Excellent agreement with the experiment is achieved when a dominant exchange interaction (|J|/k(B) approximately 73 K), along the monoclinic b axis and a sizable easy-axis magnetic anisotropy (|D|/k(B) approximately 3 K) are considered. Despite earlier suggestions for two-dimensional spin interactions, the dynamics illustrate strongly coupled antiferromagnetic S=<span class="hlt">2</span> chains and cancellation of the interchain exchange due to the lattice topology. alpha-<span class="hlt">Na</span>MnO<span class="hlt">2</span> therefore represents a model system where the geometric frustration is resolved through the lowering of the dimensionality of the spin interactions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22395945-microstructural-strength-improvements-through-use-na-sub-co-sub-cementless-ca-oh-sub-activated-class-fly-ash-system','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22395945-microstructural-strength-improvements-through-use-na-sub-co-sub-cementless-ca-oh-sub-activated-class-fly-ash-system"><span>Microstructural and strength improvements through the use of <span class="hlt">Na</span>{sub <span class="hlt">2</span>}CO{sub 3} in a cementless Ca(OH){sub <span class="hlt">2</span>}-activated Class F fly ash system</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Jeon, Dongho; Jun, Yubin; Jeong, Yeonung</p> <p>2015-01-15</p> <p>This study explores the beneficial effects of <span class="hlt">Na</span>{sub <span class="hlt">2</span>}CO{sub 3} as an additive for microstructural and strength improvements in a Ca(OH){sub <span class="hlt">2</span>}-activated fly ash system. <span class="hlt">Na</span>OH-activated fly ash samples were also tested to compare the effect of <span class="hlt">Na</span>{sub <span class="hlt">2</span>}CO{sub 3}. Compressive strength testing, XRD, SEM/BSE/EDS, {sup 29}Si/{sup 27}Al MAS-NMR, MIP and TGA were performed. The testing results indicate that the use of <span class="hlt">Na</span>{sub <span class="hlt">2</span>}CO{sub 3} for Ca(OH){sub <span class="hlt">2</span>}-activation led to a noticeable improvement in strength and microstructure, primarily due to (1) more dissolution of raw fly ash at an early age, (<span class="hlt">2</span>) more formation of C–S–H [or C–S–H(I)], (3) porositymore » reduction, and (4) pore-size refinement. We also found that (1) an early high alkalinity from the <span class="hlt">Na</span>OH formation was not a major cause of strength, (<span class="hlt">2</span>) geopolymer was not formed despite the early <span class="hlt">Na</span>OH formation, and (3) no visible pore-filling action of CaCO{sub 3} was observed. However, <span class="hlt">Na</span>{sub <span class="hlt">2</span>}CO{sub 3} did not produce any improvement in strength for <span class="hlt">Na</span>OH-activated fly ash. -- Highlights: •The use of <span class="hlt">Na</span>{sub <span class="hlt">2</span>}CO{sub 3} significantly improved strength and microstructure. •The use of <span class="hlt">Na</span>{sub <span class="hlt">2</span>}CO{sub 3} induced more dissolution of raw fly ash at early ages. •The use of <span class="hlt">Na</span>{sub <span class="hlt">2</span>}CO{sub 3} promoted more C–S–H [or C–S–H(I)] formation. •The use of <span class="hlt">Na</span>{sub <span class="hlt">2</span>}CO{sub 3} reduced total porosity and refined pore-size distribution. •The use of <span class="hlt">Na</span>{sub <span class="hlt">2</span>}CO{sub 3} produced neither geopolymer formations nor pore-filling actions from CaCO{sub 3}.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23421296','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23421296"><span>Nanocomposite dielectrics in PbO-BaO-<span class="hlt">Na</span><span class="hlt">2</span>O-Nb<span class="hlt">2</span>O5-SiO<span class="hlt">2</span> system with high breakdown strength for high voltage capacitor applications.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zhang, Qingmeng; Luo, Jun; Tang, Qun; Han, Dongfang; Zhou, Yi; Du, Jun</p> <p>2012-11-01</p> <p>Nanocomposite dielectrics in 6PbO-4BaO-20<span class="hlt">Na</span><span class="hlt">2</span>O-40Nb<span class="hlt">2</span>O5-30SiO<span class="hlt">2</span> system were prepared via melt-quenching followed by controlled crystallization. X-ray diffraction studies reveal that Pb<span class="hlt">2</span>Nb<span class="hlt">2</span>O7, Ba,<span class="hlt">Na</span>Nb5O15, <span class="hlt">Na</span>NbO3 and PbNb<span class="hlt">2</span>O6 phases are formed from the as-quenched glass annealed in temperature range from 700 degrees C to 850 degrees C. Ba<span class="hlt">2</span><span class="hlt">Na</span>Nb5O15, Pb<span class="hlt">2</span>Nb<span class="hlt">2</span>O7 crystallizes at 700 degrees C and then Pb<span class="hlt">2</span>Nb<span class="hlt">2</span>O7 disappears at 850 degrees C, while PbNb<span class="hlt">2</span>O6 and <span class="hlt">Na</span>NbO3 are formed at 850 degrees C. Microstructural observation shows that the crystallized particles are nanometer-sized and randomly distributed with glass matrix being often found at grain boundaries. The dielectric constant of the nanocomposites formed at different crystallization temperatures shows good frequency and electric field stability. The breakdown strength is slightly decreased when the glass-ceramics thickness is varied from 1 mm to 4 mm. The corresponding energy density could reach <span class="hlt">2</span>.96 J/cm3 with a breakdown strength of 58 kV/mm for thickness of 1 mm.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23860573','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23860573"><span><span class="hlt">Na</span><span class="hlt">2</span>Ti6O13: a potential anode for grid-storage sodium-ion batteries.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Rudola, Ashish; Saravanan, Kuppan; Devaraj, Sappani; Gong, Hao; Balaya, Palani</p> <p>2013-08-28</p> <p>The ultra-fast (30C or <span class="hlt">2</span> min) rate capability and impressive long cycle life (>5000 cycles) of <span class="hlt">Na</span><span class="hlt">2</span>Ti6O13 are reported. A stable <span class="hlt">2</span>.5 V sodium-ion battery full cell is demonstrated. In addition, the sodium storage mechanism and thermal stability of <span class="hlt">Na</span><span class="hlt">2</span>Ti6O13 are discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2007ApPhL..90e2911G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2007ApPhL..90e2911G"><span>Effect of (Li,Ce) doping in Aurivillius phase material <span class="hlt">Na</span>0.25K0.25Bi<span class="hlt">2</span>.5Nb<span class="hlt">2</span>O9</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gai, Zhi-Gang; Wang, Jin-Feng; Wang, Chun-Ming</p> <p>2007-01-01</p> <p>The effect of (Li,Ce) substitution for A site on the properties of <span class="hlt">Na</span>0.25K0.25Bi<span class="hlt">2</span>.5Nb<span class="hlt">2</span>O9-based ceramics was investigated. The piezoelectric activity of <span class="hlt">Na</span>0.25K0.25Bi<span class="hlt">2</span>.5Nb<span class="hlt">2</span>O9-based ceramics is significantly improved by the modification of lithium and cerium. The Curie temperature (TC) gradually increases from 668to684°C with increasing the (Li,Ce) modification. The piezoelectric coefficient d33 of the [(<span class="hlt">Na</span>0.5K0.5)Bi]0.44(LiCe)0.03[]0.03Bi<span class="hlt">2</span>Nb<span class="hlt">2</span>O9 ceramic was found to be 28pC/N, the highest value among the <span class="hlt">Na</span>0.25K0.25Bi<span class="hlt">2</span>.5Nb<span class="hlt">2</span>O9-based ceramics and also almost 50% higher than the reported d33 values of other bismuth layer-structured ferroelectric systems (˜5-19pC/N). The planar coupling factors kp and kt were found to be 8.0% and 23.0%, together with the high TC (˜670°C) and stable piezoelectric properties, demonstrating that the (Li,Ce) modified <span class="hlt">Na</span>0.25K0.25Bi<span class="hlt">2</span>.5Nb<span class="hlt">2</span>O9-based material a promising candidate for high temperature applications.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017RSOS....470921H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017RSOS....470921H"><span>Synthesis of zeolites <span class="hlt">Na</span>-A and <span class="hlt">Na</span>-X from tablet compressed and calcinated coal fly ash</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hu, Tao; Gao, Wenyan; Liu, Xin; Zhang, Yifu; Meng, Changgong</p> <p>2017-10-01</p> <p>Zeolites <span class="hlt">Na</span>-A and <span class="hlt">Na</span>-X are important synthetic zeolites widely used for separation and adsorption in industry. It is of great significance to develop energy-efficient routines that can synthesize zeolites <span class="hlt">Na</span>-A and <span class="hlt">Na</span>-X from low-cost raw materials. Coal fly ash (CFA) is the major residue from the combustion of coal and biomass containing more than 85% SiO<span class="hlt">2</span> and Al<span class="hlt">2</span>O3, which can readily replace the conventionally used sodium silicate and aluminate for zeolite synthesis. We used <span class="hlt">Na</span><span class="hlt">2</span>CO3 to replace the expensive <span class="hlt">Na</span>OH used for the calcination of CFA and showed that tablet compression can enhance the contact with <span class="hlt">Na</span><span class="hlt">2</span>CO3 for the activation of CFA through calcination for the synthesis of zeolites <span class="hlt">Na</span>-A and <span class="hlt">Na</span>-X under mild conditions. We optimized the control variables for zeolite synthesis and showed that phase-pure zeolite <span class="hlt">Na</span>-A can be synthesized with CFA at reactant molar ratio, hydrothermal reaction temperature and reaction time of 1.3<span class="hlt">Na</span><span class="hlt">2</span>O: 0.6Al<span class="hlt">2</span>O3: 1SiO<span class="hlt">2</span>: 38H<span class="hlt">2</span>O at 80°C for 6 h, respectively, while phase-pure zeolite <span class="hlt">Na</span>-X can be synthesized at <span class="hlt">2.2</span><span class="hlt">Na</span><span class="hlt">2</span>O: 0.<span class="hlt">2</span>Al<span class="hlt">2</span>O3: 1SiO<span class="hlt">2</span>: 88H<span class="hlt">2</span>O at 100°C for 8 h, respectively. The composition, morphology, specific surface area, vibration spectrum and thermogravimetry of synthesized <span class="hlt">Na</span>-A and <span class="hlt">Na</span>-X were further characterized.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/11351739','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/11351739"><span>Cesium and strontium ion exchange on the framework titanium silicate M<span class="hlt">2</span>Ti<span class="hlt">2</span>O3SiO4.nH<span class="hlt">2</span>O (M = H, <span class="hlt">Na</span>).</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Solbrå, S; Allison, N; Waite, S; Mikhalovsky, S V; Bortun, A I; Bortun, L N; Clearfield, A</p> <p>2001-02-01</p> <p>The ion exchange properties of the titanium silicate, M<span class="hlt">2</span>Ti<span class="hlt">2</span>O3SiO4.nH<span class="hlt">2</span>O (M = H, <span class="hlt">Na</span>), toward stable and radioactive 137Cs+ and 89Sr<span class="hlt">2</span>+, have been examined. By studying the cesium and strontium uptake in the presence of <span class="hlt">Na</span>NO3, CaCl<span class="hlt">2</span>, <span class="hlt">Na</span>OH, and HNO3 (in the range of 0.01-6 M) the sodium titanium silicate was found to be an efficient Cs+ ion exchanger in acid, neutral, and alkaline media and an efficient Sr<span class="hlt">2</span>+ ion exchanger in neutral and alkaline media, which makes it promising for treatment of contaminated environmental media and biological systems.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22658034-investigation-effect-structure-oxoanion-doping-na-sub-so-sub-sub-sub','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22658034-investigation-effect-structure-oxoanion-doping-na-sub-so-sub-sub-sub"><span>Investigation into the effect on structure of oxoanion doping in <span class="hlt">Na</span>{sub <span class="hlt">2</span>}M(SO{sub 4}){sub <span class="hlt">2</span>}·<span class="hlt">2</span>H{sub <span class="hlt">2</span>}O</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>Driscoll, L.L.; Kendrick, E.; Sharp Laboratories Europe, Oxford Science Park, Edmund Halley Road, Oxford OX4 4GB</p> <p>2016-10-15</p> <p>In this paper an investigation into the effect of transition metal ion and selenate/fluorophosphate doping on the structures of <span class="hlt">Na</span>{sub <span class="hlt">2</span>}M(SO{sub 4}){sub <span class="hlt">2</span>}·<span class="hlt">2</span>H{sub <span class="hlt">2</span>}O (M=transition metal) materials is reported. In agreement with previous reports, the monoclinic (Kröhnkite) structure is adopted for M=Mn, Fe, Co, Cu, while for the smallest first row divalent transition metal ion, M=Ni, the triclinic (Fairfieldite structure) is adopted. On selenate doping there is a changeover in structure from monoclinic to triclinic for M=Fe, Co, Cu, with the larger Fe{sup <span class="hlt">2</span>+} system requiring the highest level of selenate to complete the changeover. Thus the results suggest thatmore » the relative stability of the two structure types is influenced by the relative size of the transition metal: oxoanion group, with the triclinic structure favoured for small transition metals/large oxoanions. The successful synthesis of fluorophosphate doped samples, <span class="hlt">Na</span>{sub <span class="hlt">2</span>}M(SO{sub 4}){sub <span class="hlt">2</span>−x}(PO{sub 3}F){sub x}·<span class="hlt">2</span>H{sub <span class="hlt">2</span>}O was also obtained for M=Fe, Co, Cu, with the results showing a changeover in structure from monoclinic to triclinic for M=Co, Cu for very low levels (x=0.1) of fluorophosphate. In the case of M=Fe, the successful synthesis of fluorophosphates samples was achieved for x≤0.3, although no change in cell symmetry was observed. Rather in this particular case, the X-ray diffraction patterns showed evidence for selective peak broadening, attributed to local disorder as a result of the fluorophosphate group disrupting the H-bonding network. Overall the work highlights how isovalent doping can be exploited to alter the structures of <span class="hlt">Na</span>{sub <span class="hlt">2</span>}M(SO{sub 4}){sub <span class="hlt">2</span>}·<span class="hlt">2</span>H{sub <span class="hlt">2</span>}O systems. - Graphical abstract: Partial substitution of sulfate in <span class="hlt">Na</span>{sub <span class="hlt">2</span>}M(SO{sub 4}){sub <span class="hlt">2}0.2</span>H{sub <span class="hlt">2</span>}O (M=Co, Cu) by selenate or fluorophosphate leads to a structural change from the monoclinic Kröhnkite to the triclinic Fairfieldite structure. - Highlights: • The successful synthesis of <span class="hlt">Na</span>{sub <span class="hlt">2</span>}M</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2216206','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2216206"><span>Effects of muscarinic, alpha-adrenergic, and substance P agonists and ionomycin on ion transport mechanisms in the rat parotid acinar cell. The dependence of ion transport on intracellular calcium</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p></p> <p>1989-01-01</p> <p>The relationship between receptor-mediated increases in the intracellular free calcium concentration [( <span class="hlt">Ca]i</span>) and the stimulation of ion fluxes involved in fluid secretion was examined in the rat parotid acinar cell. Agonist-induced increases in [<span class="hlt">Ca]i</span> caused the rapid net loss of up to 50-60% of the total content of intracellular chloride (Cli) and potassium (Ki), which is consistent with the activation of calcium-sensitive chloride and potassium channels. These ion movements were accompanied by a 25% reduction in the intracellular volume. The relative magnitudes of the losses of Ki and the net potassium fluxes promoted by carbachol (a muscarinic agonist), phenylephrine (an alpha-adrenergic agonist), and substance P were very similar to their characteristic effects on elevating [<span class="hlt">Ca]i</span>. Carbachol stimulated the loss of Ki through multiple efflux pathways, including the large-conductance Ca-activated K channel. Carbachol and substance P increased the levels of intracellular sodium (Nai) to more than <span class="hlt">2</span>.5 times the normal level by stimulating the net uptake of sodium through multiple pathways; <span class="hlt">Na</span>-K-<span class="hlt">2</span>Cl cotransport accounted for greater than 50% of the influx, and approximately 20% was via <span class="hlt">Na</span>-H exchange, which led to a net alkalinization of the cells. Ionomycin stimulated similar fluxes through these two pathways, but also promoted sodium influx through an additional pathway which was nearly equivalent in magnitude to the combined uptake through the other two pathways. The carbachol- induced increase in Nai and decrease in Ki stimulated the activity of the sodium pump, measured by the ouabain-sensitive rate of oxygen consumption, to nearly maximal levels. In the absence of extracellular calcium or in cells loaded with the calcium chelator BAPTA (bis[o- aminophenoxy]ethane-N,N,N',N'-tetraacetic acid) the magnitudes of agonist- or ionomycin-stimulated ion fluxes were greatly reduced. The parotid cells displayed a marked desensitization to substance P; within 10 min the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017ECSS..198..450K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017ECSS..198..450K"><span>Abundance, distribution and bioavailability of major and trace elements in surface sediments from the <span class="hlt">Cai</span> River estuary and Nha Trang Bay (South China Sea, Vietnam)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Koukina, S. E.; Lobus, N. V.; Peresypkin, V. I.; Dara, O. M.; Smurov, A. V.</p> <p>2017-11-01</p> <p>Major (Si, Al, Fe, Ti, Mg, Ca, <span class="hlt">Na</span>, K, S, P), minor (Mn) and trace (Li, V, Cr, Co, Ni, Cu, Zn, As, Sr, Zr, Mo, Cd, Ag, Sn, Sb, Cs, Ba, Hg, Pb, Bi and U) elements, their chemical forms and the mineral composition, organic matter (TOC) and carbonates (TIC) in surface sediments from the <span class="hlt">Cai</span> River estuary and Nha Trang Bay were first determined along the salinity gradient. The abundance and ratio of major and trace elements in surface sediments are discussed in relation to the mineralogy, grain size, depositional conditions, reference background and SQG values. Most trace-element contents are at natural levels and are derived from the composition of rocks and soils in the watershed. A severe enrichment of Ag is most likely derived from metal-rich detrital heavy minerals such as Ag-sulfosalts. Along the salinity gradient, several zones of metal enrichment occur in surface sediments because of the geochemical fractionation of the riverine material. The parts of actually and potentially bioavailable forms (isolated by four single chemical reagent extractions) are most elevated for Mn and Pb (up to 36% and 32% of total content, respectively). The possible anthropogenic input of Pb in the region requires further study. Overall, the most bioavailable parts of trace elements are associated with easily soluble amorphous Fe and Mn oxyhydroxides. The sediments are primarily enriched with bioavailable metal forms in the riverine part of the estuary. Natural (such as turbidities) and human-generated (such as urban and industrial activities) pressures are shown to influence the abundance and speciation of potential contaminants and therefore change their bioavailability in this estuarine system.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4170523','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4170523"><span>Structural analyses of Ca<span class="hlt">2</span>+/CaM interaction with <span class="hlt">Na</span>V channel C-termini reveal mechanisms of calcium-dependent regulation</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Wang, Chaojian; Chung, Ben C.; Yan, Haidun; Wang, Hong-Gang; Lee, Seok-Yong; Pitt, Geoffrey S.</p> <p>2014-01-01</p> <p>Ca<span class="hlt">2</span>+ regulates voltage-gated <span class="hlt">Na</span>+ (<span class="hlt">Na</span>V) channels and perturbed Ca<span class="hlt">2</span>+ regulation of <span class="hlt">Na</span>V function is associated with epilepsy syndromes, autism, and cardiac arrhythmias. Understanding the disease mechanisms, however, has been hindered by a lack of structural information and competing models for how Ca<span class="hlt">2</span>+ affects <span class="hlt">Na</span>V channel function. Here, we report the crystal structures of two ternary complexes of a human <span class="hlt">Na</span>V cytosolic C-terminal domain (CTD), a fibroblast growth factor homologous factor, and Ca<span class="hlt">2</span>+/calmodulin (Ca<span class="hlt">2</span>+/CaM). These structures rule out direct binding of Ca<span class="hlt">2</span>+ to the <span class="hlt">Na</span>V CTD, and uncover new contacts between CaM and the <span class="hlt">Na</span>V CTD. Probing these new contacts with biochemical and functional experiments allows us to propose a mechanism by which Ca<span class="hlt">2</span>+ could regulate <span class="hlt">Na</span>V channels. Further, our model provides hints towards understanding the molecular basis of the neurologic disorders and cardiac arrhythmias caused by <span class="hlt">Na</span>V channel mutations. PMID:25232683</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19900018593','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19900018593"><span>Evaluation of EA-934<span class="hlt">NA</span> with <span class="hlt">2</span>.5 percent Cab-O-Sil</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Caldwell, Gordon A.</p> <p>1990-01-01</p> <p>Currently, Hysol adhesive EA-934<span class="hlt">NA</span> is used to bond the Field Joint Protection System on the Shuttle rocket motors at Kennedy Space Center. However, due to processing problems, an adhesive with a higher viscosity is needed to alleviate these difficulties. One possible solution is to add Cab-O-Sil to the current adhesive. The adhesive strength and bond strengths that can be obtained when <span class="hlt">2</span>.5 percent Cab-O-Sil is added to adhesive EA-934<span class="hlt">NA</span> are examined and tested over a range of test temperatures from -20 to 300 F. Tensile adhesion button and lap shear specimens were bonded to D6AC steel and uniaxial tensile specimens (testing for strength, initial tangent modulus, elongation and Poisson's ratio) were prepared using Hysol adhesive EA-934<span class="hlt">NA</span> with <span class="hlt">2</span>.5 percent Cab-O-Sil added. These specimens were tested at -20, 20, 75, 100, 125, 150, 200, 250, and 300 F, respectively. Additional tensile adhesion button specimens bonding Rust-Oleum primed and painted D6AC steel to itself and to cork using adhesive EA-934<span class="hlt">NA</span> with <span class="hlt">2</span>.5 percent Cab-O-Sil added were tested at 20, 75, 125, 200, and 300 F, respectively. Results generally show decreasing strength values with increasing test temperatures. The bond strengths obtained using cork as a substrate were totally dependent on the cohesive strength of the cork.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27327414','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27327414"><span>Core-shell-shell heterostructures of α-<span class="hlt">Na</span>LuF4:Yb/Er@<span class="hlt">Na</span>LuF4:Yb@MF<span class="hlt">2</span> (M = Ca, Sr, Ba) with remarkably enhanced upconversion luminescence.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Su, Yue; Liu, Xiuling; Lei, Pengpeng; Xu, Xia; Dong, Lile; Guo, Xianmin; Yan, Xingxu; Wang, Peng; Song, Shuyan; Feng, Jing; Zhang, Hongjie</p> <p>2016-07-05</p> <p>Core-shell-shell heterostructures of α-<span class="hlt">Na</span>LuF4:Yb/Er@<span class="hlt">Na</span>LuF4:Yb@MF<span class="hlt">2</span> (M = Ca, Sr, Ba) have been successfully fabricated via the thermal decomposition method. Upconversion nanoparticles (UCNPs) were characterized by powder X-ray diffraction (XRD), transmission electron microscopy (TEM), upconversion luminescence (UCL) spectroscopy, etc. Under 980 nm excitation, the emission intensities of the UCNPs are remarkably enhanced after coating the MF<span class="hlt">2</span> (M = Ca, Sr, and Ba) shell. Among these samples, CaF<span class="hlt">2</span> coated UCNPs show the strongest overall emission, while BaF<span class="hlt">2</span> coated UCNPs exhibit the longest lifetime. These results demonstrate that alkaline earth metal fluorides are ideal materials to improve the UCL properties. Meanwhile, although the lattice mismatch between the ternary <span class="hlt">Na</span>REF4 core and the binary MF<span class="hlt">2</span> (M = Sr and Ba) shell is relatively large, the successfully synthesized <span class="hlt">Na</span>LuF4:Yb/Er@<span class="hlt">Na</span>LuF4:Yb@MF<span class="hlt">2</span> indicates a new outlook on the fabrication of heterostructural core-shell UCNPs.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2001JPS....97..465R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2001JPS....97..465R"><span>Electrochemical properties and structures of the mixed-valence lithium cuprates Li 3Cu <span class="hlt">2</span>O 4 and Li <span class="hlt">2</span><span class="hlt">Na</span>Cu <span class="hlt">2</span>O 4</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Raekelboom, E. A.; Hector, A. L.; Weller, M. T.; Owen, J. R.</p> <p></p> <p>The electrochemical performances of Li 3Cu <span class="hlt">2</span>O 4 and Li <span class="hlt">2</span><span class="hlt">Na</span>Cu <span class="hlt">2</span>O 4 as cathode materials in lithium coin type batteries have been studied. In Li 3Cu <span class="hlt">2</span>O 4, the copper was oxidised to the III level when cycling. The replacement of the lithium by the sodium ions in the octahedral sites in Li <span class="hlt">2</span><span class="hlt">Na</span>Cu <span class="hlt">2</span>O 4 might have an effect on the pathway of the lithium ions during the (de)intercalations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JPCS..102...34N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JPCS..102...34N"><span><span class="hlt">Na</span>F-assisted combustion synthesis of MoSi<span class="hlt">2</span> nanoparticles and their densification behavior</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Nersisyan, Hayk H.; Lee, Tae Hyuk; Ri, Vladislav; Lee, Jong Hyeon; Suh, Hoyoung; Kim, Jin-Gyu; Son, Hyeon Taek; Kim, Yong-Ho</p> <p>2017-03-01</p> <p>The exothermic reduction of oxides mixture (MoO3+<span class="hlt">2</span>SiO<span class="hlt">2</span>) by magnesium in <span class="hlt">Na</span>F melt enables the synthesis of nanocrystalline MoSi<span class="hlt">2</span> powders in near-quantitative yields. The combustion wave with temperature of about 1000-1200 °C was recorded in highly diluted by <span class="hlt">Na</span>F starting mixtures. The by-products of combustion reaction (<span class="hlt">Na</span>F and MgO) were subsequently removed by leaching with acid and washing with water. The as-prepared MoSi<span class="hlt">2</span> nanopowder composed of spherical and dendritic shape particles was consolidated using the spark plasma sintering method at 1200-1500 °C and 50 MPa for 10 min. The result was dense compacts (98.6% theoretical density) possessing submicron grains and exhibiting hardness of 8.74-12.92 GPa.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29790854','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29790854"><span>Superior ionic and electronic properties of ReN<span class="hlt">2</span> monolayers for <span class="hlt">Na</span>-ion battery electrodes.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zhang, Shi-Hao; Liu, Bang-Gui</p> <p>2018-08-10</p> <p>Excellent monolayer electrode materials can be used to design high-performance alkali-metal-ion batteries. Here, we propose two-dimensional ReN <span class="hlt">2</span> monolayers as superior sodium-ion battery materials. Our total energy optimization results in a buckled tetragonal structure for the ReN <span class="hlt">2</span> monolayer, and our phonon spectrum and elastic moduli prove that it is dynamically and mechanically stable. Further investigations show that it is metallic and still keeps its metallic feature after the adsorption of <span class="hlt">Na</span> or K atoms, and the adsorption of <span class="hlt">Na</span> (or K) atoms changes the lattice parameters by 3.<span class="hlt">2</span>% (or 3.8%) at most. Its maximum capacity reaches 751 mA h g -1 for <span class="hlt">Na</span>-ion batteries or 250 mA h g -1 for K-ion batteries, and the diffusion barrier is only 0.027 eV for the <span class="hlt">Na</span> atom or 0.127 eV for the K atom. The small lattice changes, high storage capacity, metallic feature, and extremely low ion diffusion barriers make the ReN <span class="hlt">2</span> monolayers a superior electrode material for <span class="hlt">Na</span>-ion rechargeable batteries with ultrafast charging/discharging processes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22658297-synthesis-structural-characterization-hexagonal-anti-perovskite-na-sub-cavo-sub','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22658297-synthesis-structural-characterization-hexagonal-anti-perovskite-na-sub-cavo-sub"><span>Synthesis and structural characterization of the hexagonal anti-perovskite <span class="hlt">Na</span>{sub <span class="hlt">2</span>}CaVO{sub 4}F</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>Green, Robert L., E-mail: rgreen@flpoly.org; Avdeev, Maxim; School of Chemistry, The University of Sydney, Sydney, NSW 2006</p> <p></p> <p>The structural details of the ordered hexagonal oxyfluoride <span class="hlt">Na</span>{sub <span class="hlt">2</span>}CaVO{sub 4}F prepared by solid-state synthesis using stoichiometric amounts of V{sub <span class="hlt">2</span>}O{sub 5}, CaCO{sub 3}, <span class="hlt">Na</span>{sub <span class="hlt">2</span>}CO{sub 3} and <span class="hlt">Na</span>F were characterized using high-resolution neutron powder diffraction. The structural changes between 25 °C and 750 °C revealed that the two structural subunits in this material behave different when heated: there is an expansion of the face-shared FNa{sub 4}Ca{sub <span class="hlt">2</span>} octahedra while the VO{sub 4} tetrahedra due to increased thermal disorder reveal marginal bond contractions. Bond valences and the global instability index point to significant structural disorder at 750 °C. - Graphicalmore » abstract: The structure of the novel oxyfluoride <span class="hlt">Na</span>{sub <span class="hlt">2</span>}CaVO{sub 4}F is studied at room temperature and high-temperatures. The structure can be viewed as layers of compression and elongation of polyhedral subunits, which change as a function of temperature. - Highlights: • The novel oxyfluoride, <span class="hlt">Na</span>{sub <span class="hlt">2</span>}CaVO{sub 4}F, is synthesized via solid-state method. • High-resolution neutron diffraction data is used to analyze the structure of <span class="hlt">Na</span>{sub <span class="hlt">2</span>}CaVO{sub 4}F. • Structural subunits exhibit expansion and contraction with increasing temperature. • Higher temperatures increase instability within the structure of <span class="hlt">Na</span>{sub <span class="hlt">2</span>}CaVO{sub 4}F.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JMMM..458...62S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JMMM..458...62S"><span>Magnetic properties of superparamagnetic β-<span class="hlt">Na</span>FeO<span class="hlt">2</span> nanoparticles</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Singh, Sarbjit; Tovstolytkin, Alexandr; Lotey, Gurmeet Singh</p> <p>2018-07-01</p> <p>Superparamagnetic β-<span class="hlt">Na</span>FeO<span class="hlt">2</span> nanoparticles of particle size 37 nm with orthorhombic crystal structure and space group Pn21a have been prepared by sol-gel method. Temperature dependent magnetic study has been performed. Its systematic analysis has been done to calculate the Curie and blocking temperatures along with its magnetic susceptibility. The Langevin fitting of the magnetic data has been carried out. It has been shown that the synthesized nanoparticles exhibit superparamagnetic behavior. The Neel's relaxation time has been calculated to further support its superparamagnetic nature. The synthesized β-<span class="hlt">Na</span>FeO<span class="hlt">2</span> nanoparticles behave like ferromagnets below 80 K; they are superparamagnetic above 80 K-340 K and thereafter as paramagnetic. The possible mechanism of superparamagnetism has been discussed. It has been concluded that these nanoparticles can find wide applications in the area of biomedical sciences.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li class="active"><span>22</span></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li><a href="#" onclick='return showDiv("page_24");'>24</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_22 --> <div id="page_23" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li class="active"><span>23</span></li> <li><a href="#" onclick='return showDiv("page_24");'>24</a></li> <li><a href="#" onclick='return showDiv("page_25");'>25</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="441"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19830025745','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19830025745"><span>Effect of the amount of <span class="hlt">Na</span><span class="hlt">2</span>SO4 on the high temperature corrosion of Udimet-700</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Misra, A. K.; Kohl, F. J.</p> <p>1983-01-01</p> <p>The corrosion of Udimet-700, coated with different doses of <span class="hlt">Na</span><span class="hlt">2</span>SO4, was studied in an isothermal thermogravimetric test in the temperature range 900 to 950 C. The weight gain curve is characterized by five distinct stages: an initial period of linear corrosion; an induction period; a period of accelerated corrosion; a period of decelerating corrosion; and a period of parabolic oxidation. The time required for the failure of the alloy increases with an increase in the amount of <span class="hlt">Na</span><span class="hlt">2</span>SO4, reaches a peak and then decreases with further increase in the amount of <span class="hlt">Na</span><span class="hlt">2</span>SO4. For low and intermediate doses (0.3 to <span class="hlt">2</span>.0 mg/sq cm), the catastrophic failure of the material occurs by the formation of <span class="hlt">Na</span><span class="hlt">2</span>MoO4 and interaction of the liquid <span class="hlt">Na</span><span class="hlt">2</span>MoO4 with the alloy. For heavy doses, the degradation of the material is due to the formation of large amounts of sulfides.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015JaJAP..54hKC20M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015JaJAP..54hKC20M"><span>First-principles study on alkali-metal effect of Li, <span class="hlt">Na</span>, and K in CuInSe<span class="hlt">2</span> and CuGaSe<span class="hlt">2</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Maeda, Tsuyoshi; Kawabata, Atsuhito; Wada, Takahiro</p> <p>2015-08-01</p> <p>The substitution energies and migration energies of the alkali metal atoms of Li, <span class="hlt">Na</span>, and K in CuInSe<span class="hlt">2</span> (CIS) and CuGaSe<span class="hlt">2</span> (CGS) were investigated by first-principles calculations. The substitution energies of Li, <span class="hlt">Na</span>, and K atoms in CIS and CGS were calculated for two different cationic atom positions of Cu and In/Ga in the chalcopyrite unit cell. In CIS and CGS, the substitution energies of <span class="hlt">Na</span>Cu are much lower than those of <span class="hlt">Na</span>In and <span class="hlt">Na</span>Ga. The substitution energies of the LiCu atoms in CIS and CGS are lower than those of <span class="hlt">Na</span>Cu, while the substitution energies of KCu atoms in CIS and CGS are much higher than those of <span class="hlt">Na</span>Cu. Therefore, it is difficult to form KCu in CIS and CGS. The migration energies of Li, <span class="hlt">Na</span>, and K atoms in CIS and CGS are obtained by a combination of the linear and quadratic synchronous transit (LST/QST) methods and the nudged elastic band (NEB) method. The theoretical migration energies of a <span class="hlt">Na</span> atom at the Cu site to the nearest Cu vacancy (<span class="hlt">Na</span>Cu → VCu) in CIS and CGS are much lower than those of (CuCu → VCu) in CIS and CGS. The mechanism underlying the alkali metal effect of Li, <span class="hlt">Na</span>, and K in the CIGS film during the post-deposition treatment of LiF, <span class="hlt">Na</span>F, and KF is discussed on the basis of the calculated substitution and migration energies.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29461536','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29461536"><span>Oxygen redox chemistry without excess alkali-metal ions in <span class="hlt">Na</span><span class="hlt">2</span>/3[Mg0.28Mn0.72]O<span class="hlt">2</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Maitra, Urmimala; House, Robert A; Somerville, James W; Tapia-Ruiz, Nuria; Lozano, Juan G; Guerrini, Niccoló; Hao, Rong; Luo, Kun; Jin, Liyu; Pérez-Osorio, Miguel A; Massel, Felix; Pickup, David M; Ramos, Silvia; Lu, Xingye; McNally, Daniel E; Chadwick, Alan V; Giustino, Feliciano; Schmitt, Thorsten; Duda, Laurent C; Roberts, Matthew R; Bruce, Peter G</p> <p>2018-03-01</p> <p>The search for improved energy-storage materials has revealed Li- and <span class="hlt">Na</span>-rich intercalation compounds as promising high-capacity cathodes. They exhibit capacities in excess of what would be expected from alkali-ion removal/reinsertion and charge compensation by transition-metal (TM) ions. The additional capacity is provided through charge compensation by oxygen redox chemistry and some oxygen loss. It has been reported previously that oxygen redox occurs in O <span class="hlt">2</span>p orbitals that interact with alkali ions in the TM and alkali-ion layers (that is, oxygen redox occurs in compounds containing Li + -O(<span class="hlt">2</span>p)-Li + interactions). <span class="hlt">Na</span> <span class="hlt">2</span>/3 [Mg 0.28 Mn 0.72 ]O <span class="hlt">2</span> exhibits an excess capacity and here we show that this is caused by oxygen redox, even though Mg <span class="hlt">2</span>+ resides in the TM layers rather than alkali-metal (AM) ions, which demonstrates that excess AM ions are not required to activate oxygen redox. We also show that, unlike the alkali-rich compounds, <span class="hlt">Na</span> <span class="hlt">2</span>/3 [Mg 0.28 Mn 0.72 ]O <span class="hlt">2</span> does not lose oxygen. The extraction of alkali ions from the alkali and TM layers in the alkali-rich compounds results in severely underbonded oxygen, which promotes oxygen loss, whereas Mg <span class="hlt">2</span>+ remains in <span class="hlt">Na</span> <span class="hlt">2</span>/3 [Mg 0.28 Mn 0.72 ]O <span class="hlt">2</span> , which stabilizes oxygen.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018NatCh..10..288M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018NatCh..10..288M"><span>Oxygen redox chemistry without excess alkali-metal ions in <span class="hlt">Na</span><span class="hlt">2</span>/3[Mg0.28Mn0.72]O<span class="hlt">2</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Maitra, Urmimala; House, Robert A.; Somerville, James W.; Tapia-Ruiz, Nuria; Lozano, Juan G.; Guerrini, Niccoló; Hao, Rong; Luo, Kun; Jin, Liyu; Pérez-Osorio, Miguel A.; Massel, Felix; Pickup, David M.; Ramos, Silvia; Lu, Xingye; McNally, Daniel E.; Chadwick, Alan V.; Giustino, Feliciano; Schmitt, Thorsten; Duda, Laurent C.; Roberts, Matthew R.; Bruce, Peter G.</p> <p>2018-03-01</p> <p>The search for improved energy-storage materials has revealed Li- and <span class="hlt">Na</span>-rich intercalation compounds as promising high-capacity cathodes. They exhibit capacities in excess of what would be expected from alkali-ion removal/reinsertion and charge compensation by transition-metal (TM) ions. The additional capacity is provided through charge compensation by oxygen redox chemistry and some oxygen loss. It has been reported previously that oxygen redox occurs in O <span class="hlt">2</span>p orbitals that interact with alkali ions in the TM and alkali-ion layers (that is, oxygen redox occurs in compounds containing Li+-O(<span class="hlt">2</span>p)-Li+ interactions). <span class="hlt">Na</span><span class="hlt">2</span>/3[Mg0.28Mn0.72]O<span class="hlt">2</span> exhibits an excess capacity and here we show that this is caused by oxygen redox, even though Mg<span class="hlt">2</span>+ resides in the TM layers rather than alkali-metal (AM) ions, which demonstrates that excess AM ions are not required to activate oxygen redox. We also show that, unlike the alkali-rich compounds, <span class="hlt">Na</span><span class="hlt">2</span>/3[Mg0.28Mn0.72]O<span class="hlt">2</span> does not lose oxygen. The extraction of alkali ions from the alkali and TM layers in the alkali-rich compounds results in severely underbonded oxygen, which promotes oxygen loss, whereas Mg<span class="hlt">2</span>+ remains in <span class="hlt">Na</span><span class="hlt">2</span>/3[Mg0.28Mn0.72]O<span class="hlt">2</span>, which stabilizes oxygen.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JPS...385..114L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JPS...385..114L"><span>The different Li/<span class="hlt">Na</span> ion storage mechanisms of nano Sb<span class="hlt">2</span>O3 anchored on graphene</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Li, Hai; Qian, Kun; Qin, Xianying; Liu, Dongqing; Shi, Ruiying; Ran, Aihua; Han, Cuiping; He, Yan-Bing; Kang, Feiyu; Li, Baohua</p> <p>2018-05-01</p> <p>The antimony oxide/reduced graphene oxide (Sb<span class="hlt">2</span>O3/rGO) nanocomposites are used as anode of Li-ion and <span class="hlt">Na</span>-ion batteries (LIBs and NIBs). However, it is unclear about Li-ion and <span class="hlt">Na</span>-ion storage mechanism in Sb<span class="hlt">2</span>O3/rGO nanocomposites. Herein, the conversion-alloying mechanisms of Sb<span class="hlt">2</span>O3/rGO anodes for <span class="hlt">Na</span>-ion and Li-ion storage are comparatively studied with a combined in-situ XRD and quasi in-situ XPS method. The distinct behaviours are monitored during (de)lithiation and (de)sodiation with respect to crystal structure and chemical composition evolution. It is evidenced that the <span class="hlt">Na</span>-ion can be easily transported to the inner part of the Sb<span class="hlt">2</span>O3, where the Li-ion almost cannot reach, leading to a fully transformation during sodiation. In addition, the conversion reaction product of amorphous <span class="hlt">Na</span><span class="hlt">2</span>O display their better chemical stability than amorphous Li<span class="hlt">2</span>O during electrochemical cycles, which contribute to a stable and long cycling life of NIBs. This work gain insight into the high-capacity anodes with conversation-alloying mechanism for NIBs.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2008OptMa..30..687H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2008OptMa..30..687H"><span>The electronic structure and optical properties of ABP <span class="hlt">2</span>O 7 ( A = <span class="hlt">Na</span>, Li) double phosphates</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hizhnyi, Yu. A.; Oliynyk, A.; Gomenyuk, O.; Nedilko, S. G.; Nagornyi, P.; Bojko, R.; Bojko, V.</p> <p>2008-01-01</p> <p>Partial densities of states and reflection spectra of <span class="hlt">Na</span>AlP <span class="hlt">2</span>O 7, KAlP <span class="hlt">2</span>O 7 and LiInP <span class="hlt">2</span>O 7 double phosphate crystals are calculated by the full-potential linear-augmented-plane-wave (FLAPW) method. Experimental reflection spectra of KAlP <span class="hlt">2</span>O 7, CsAlP <span class="hlt">2</span>O 7 and <span class="hlt">Na</span>InP <span class="hlt">2</span>O 7 are measured in the 4-20 eV energy range. The values of band gaps, Eg, are found from a comparison of experiment and calculations to be 6.0 eV for <span class="hlt">Na</span>AlP <span class="hlt">2</span>O 7 and KAlP <span class="hlt">2</span>O 7, and 4.6 eV for LiInP <span class="hlt">2</span>O 7.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016GeCoA.177..170T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016GeCoA.177..170T"><span>Rutile solubility in <span class="hlt">NaF-Na</span>Cl-KCl-bearing aqueous fluids at 0.5-<span class="hlt">2</span>.79 GPa and 250-650 °C</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Tanis, Elizabeth A.; Simon, Adam; Zhang, Youxue; Chow, Paul; Xiao, Yuming; Hanchar, John M.; Tschauner, Oliver; Shen, Guoyin</p> <p>2016-03-01</p> <p> during the experiment, and the measured concentration of Zr in the fluid was used to calculate the concentration of Ti (i.e., the solubility of rutile) in the fluid. The salts <span class="hlt">Na</span>F, <span class="hlt">Na</span>Cl, and KCl were systematically added to the aqueous fluid, and the relative effects of fluid composition, pressure, and temperature on rutile solubility were quantified. The results indicate that fluid composition exerts the greatest control on rutile solubility in aqueous fluid, consistent with previous studies, and that increasing temperature has a positive, albeit less pronounced, effect. The solubility of Zr-rutile in aqueous fluid increases with the addition of halides in the following order: <span class="hlt">2</span> wt% <span class="hlt">Na</span>F < 30 wt% KCl < 30 wt% <span class="hlt">Na</span>Cl < 3 wt% <span class="hlt">Na</span>F < (10 wt% <span class="hlt">Na</span>Cl + <span class="hlt">2</span> wt% <span class="hlt">Na</span>F) < 4 wt% <span class="hlt">Na</span>F. The solubility of rutile in the fluid increases with the <span class="hlt">2</span>nd to 3rd power of the Cl- concentration, and the 3rd to 4th power of the F- concentration. These new data are consistent with observations from field studies of exhumed terranes that indicate that rutile is soluble in complex aqueous fluids, and that fluid composition is the primary control on rutile solubility and HFSE mobility.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1046158','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1046158"><span>POST-OPERATIONAL TREATMENT OF RESIDUAL <span class="hlt">NA</span> COOLLANT IN EBR-<span class="hlt">2</span> USING CARBONATION</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>Sherman, S.; Knight, C.</p> <p>2011-03-08</p> <p>At the end of 2002, the Experimental Breeder Reactor Two (EBR-II) facility became a U.S. Resource Conservation and Recovery Act (RCRA) permitted site, and the RCRA permit1 compelled further treatment of the residual sodium in order to convert it into a less reactive chemical form and remove the by-products from the facility, so that a state of RCRA 'closure' for the facility may be achieved (42 U.S.C. 6901-6992k, 2002). In response to this regulatory driver, and in recognition of project budgetary and safety constraints, it was decided to treat the residual sodium in the EBR-II primary and secondary sodium systemsmore » using a process known as 'carbonation.' In early EBR-II post-operation documentation, this process is also called 'passivation.' In the carbonation process (Sherman and Henslee, 2005), the system containing residual sodium is flushed with humidified carbon dioxide (CO{sub <span class="hlt">2</span>}). The water vapor in the flush gas reacts with residual sodium to form sodium hydroxide (<span class="hlt">Na</span>OH), and the CO{sub <span class="hlt">2</span>} in the flush gas reacts with the newly formed <span class="hlt">Na</span>OH to make sodium bicarbonate (<span class="hlt">Na</span>HCO{sub 3}). Hydrogen gas (H{sub <span class="hlt">2</span>}) is produced as a by-product. The chemical reactions occur at the exposed surface of the residual sodium. The <span class="hlt">Na</span>HCO{sub 3} layer that forms is porous, and humidified carbon dioxide can penetrate the <span class="hlt">Na</span>HCO{sub 3} layer to continue reacting residual sodium underneath. The rate of reaction is controlled by the thickness of the <span class="hlt">Na</span>HCO{sub 3} surface layer, the moisture input rate, and the residual sodium exposed surface area. At the end of carbonation, approximately 780 liters of residual sodium in the EBR-II primary tank ({approx}70% of original inventory), and just under 190 liters of residual sodium in the EBR-II secondary sodium system ({approx}50% of original inventory), were converted into <span class="hlt">Na</span>HCO{sub 3}. No bare surfaces of residual sodium remained after treatment, and all remaining residual sodium deposits are covered by a layer of <span class="hlt">Na</span>HCO{sub 3</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/21494202-fluorine-sites-glasses-transparent-glass-ceramics-system-na-sub-sub-al-sub-sub-sio-sub-baf-sub','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/21494202-fluorine-sites-glasses-transparent-glass-ceramics-system-na-sub-sub-al-sub-sub-sio-sub-baf-sub"><span>Fluorine sites in glasses and transparent glass-ceramics of the system <span class="hlt">Na</span>{sub <span class="hlt">2</span>}O/K{sub <span class="hlt">2</span>}O/Al{sub <span class="hlt">2</span>}O{sub 3}/SiO{sub <span class="hlt">2</span>}/BaF{sub <span class="hlt">2</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>Bocker, Christian, E-mail: christian.bocker@uni-jena.d; Munoz, Francisco; Duran, Alicia</p> <p>2011-02-15</p> <p>The transparent glass-ceramics obtained in the silicate system <span class="hlt">Na</span>{sub <span class="hlt">2</span>}O/K{sub <span class="hlt">2</span>}O/SiO{sub <span class="hlt">2</span>}/BaF{sub <span class="hlt">2</span>} show homogeneously dispersed BaF{sub <span class="hlt">2</span>} nano crystals with a narrow size distribution. The X-ray diffraction and the nuclear magnetic resonance spectroscopy were applied to glasses and the respective glass-ceramics in order to clarify the crystallization mechanism and the role of fluorine during crystallization. With an increasing annealing time, the concentration and also the number of crystals remain approximately constant. With an increasing annealing temperature, the crystalline fraction increases until a saturation limit is reached, while the number of crystals decreases and the size of the crystals increases.more » Fluoride in the glassy network occurs as Al-F-Ba, Al-F-<span class="hlt">Na</span> and also as Ba-F structures. The latter are transformed into crystalline BaF{sub <span class="hlt">2</span>} and fluoride is removed from the Al-F-Ba/<span class="hlt">Na</span> bonds. However, some fluorine is still present in the glassy phase after the crystallization. -- Graphical abstract: The X-ray diffraction and the nuclear magnetic resonance spectroscopy were applied to glasses in the silicate system <span class="hlt">Na</span>{sub <span class="hlt">2</span>}O/K{sub <span class="hlt">2</span>}O/SiO{sub <span class="hlt">2</span>}/BaF{sub <span class="hlt">2</span>} and the respective glass-ceramics with BaF{sub <span class="hlt">2</span>} nano crystals in order to clarify the crystallization mechanism and the role of fluorine during crystallization. Display Omitted Research highlights: {yields} BaF{sub <span class="hlt">2</span>} nano crystals are precipitated from a silicate glass system. {yields} Ostwald ripening during the late stage of crystallization does not occur. {yields} Fluorine in the glass is coordinated with Ba as well as Al together with Ba or <span class="hlt">Na</span>.{yields} In the glass-ceramics, the residual fluorine is coordinated as Al-F-Ba/<span class="hlt">Na</span>.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/17371658','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/17371658"><span>Contributing factors to chronic ankle instability.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Hubbard, Tricia J; Kramer, Lauren C; Denegar, Craig R; Hertel, Jay</p> <p>2007-03-01</p> <p>The development of repetitive ankle sprains and persistent symptoms after initial ankle sprain has been termed chronic ankle instability (<span class="hlt">CAI</span>). There is no clear indication of which measures are most important in discriminating between individuals with and without <span class="hlt">CAI</span>. Thirty subjects with unilateral <span class="hlt">CAI</span> and controls had measures of ankle laxity and hypomobility, static and dynamic balance, ankle and hip strength, lower extremity alignments, and flexibility taken on both limbs. Based on comparisons of <span class="hlt">CAI</span> ankles and side-matched limbs in controls, the measures significantly predictive of <span class="hlt">CAI</span> were increased inversion laxity (r(<span class="hlt">2</span>) change = 0.203), increased anterior laxity (r(<span class="hlt">2</span>) change = 0.11), more missed balance trials (r(<span class="hlt">2</span>) change = 0.094), and lower plantarflexion to dorsiflexion peak torque (r(<span class="hlt">2</span>) change = 0.052). Symmetry indices comparing the side-to-side differences of each measure also were calculated for each dependent variable and compared between groups. The measures significantly predictive of <span class="hlt">CAI</span> were decreased anterior reach (r(<span class="hlt">2</span>) change = 0.185), decreased plantarflexion peak torque (r(<span class="hlt">2</span>) change = 0.099), decreased posterior medial reach (r(<span class="hlt">2</span>) change = 0.094), and increased inversion laxity (r(<span class="hlt">2</span>) change = 0.041). The results of this study elucidate the specific measures that best discriminate between individuals with and without <span class="hlt">CAI</span>. Both mechanical (anterior and inversion laxity) and functional (strength, dynamic balance) insufficiencies significantly contribute to the etiology of <span class="hlt">CAI</span>. Prevention of <span class="hlt">CAI</span> may be possible with proper initial management of the acute injury with rehabilitation aimed at those factors that best discriminate between individuals with and without <span class="hlt">CAI</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014PhRvB..89e4105G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014PhRvB..89e4105G"><span>Theoretical prediction of morphotropic compositions in <span class="hlt">Na</span>1/<span class="hlt">2</span>Bi1/<span class="hlt">2</span>TiO3-based solid solutions from transition pressures</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gröting, Melanie; Albe, Karsten</p> <p>2014-02-01</p> <p>In this article we present a method based on ab initio calculations to predict compositions at morphotropic phase boundaries in lead-free perovskite solid solutions. This method utilizes the concept of flat free energy surfaces and involves the monitoring of pressure-induced phase transitions as a function of composition. As model systems, solid solutions of <span class="hlt">Na</span>1/<span class="hlt">2</span>Bi1/<span class="hlt">2</span>TiO3 with the alkali substituted Li1/<span class="hlt">2</span>Bi1/<span class="hlt">2</span>TiO3 and K1/<span class="hlt">2</span>Bi1/<span class="hlt">2</span>TiO3 and the alkaline earth substituted CaTiO3 and BaTiO3 are chosen. The morphotropic compositions are identified by determining the composition at which the phase transition pressure equals zero. In addition, we discuss the different effects of hydrostatic pressure (compression and tension) and chemical substitution on the antiphase tilts about the [111] axis (a-a-a-) present in pure <span class="hlt">Na</span>1/<span class="hlt">2</span>Bi1/<span class="hlt">2</span>TiO3 and how they develop in the two solid solutions <span class="hlt">Na</span>1/<span class="hlt">2</span>Bi1/<span class="hlt">2</span>TiO3-CaTiO3 and <span class="hlt">Na</span>1/<span class="hlt">2</span>Bi1/<span class="hlt">2</span>TiO3-BaTiO3. Finally, we discuss the advantages and shortcomings of this simple computational approach.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29537696','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29537696"><span>TiO<span class="hlt">2</span> Nanostructures as Anode Materials for Li/<span class="hlt">Na</span>-ion Batteries.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Vazquez-Santos, Maria B; Tartaj, Pedro; Morales, Enrique; Amarilla, Jose Manuel</p> <p>2018-03-14</p> <p>Here we summarize some results on the use of TiO <span class="hlt">2</span> nanostructures as anode materials for more efficient Li-ion (LIBs) and <span class="hlt">Na</span>-ion (NIBs) batteries. LIBs are the leader to power portable electronic devices, and represent in the short-term the most adequate technology to power electrical vehicles, while NIBs hold promise for large storage of energy generated from renewable sources. Specifically, TiO <span class="hlt">2</span> an abundant, low cost, chemically stable and environmentally safe oxide represents in LIBs an alternative to graphite for applications in which safety is mandatory. For NIBs, TiO <span class="hlt">2</span> anodes (or more precisely negative electrodes) work at low voltage, assuring acceptable energy density values. Finally, assembling different TiO <span class="hlt">2</span> polymorphs in the form of nanostructures decreases diffusion distances, increases the number of contacts and offering additional sites for <span class="hlt">Na</span> + storage, helping to improve power efficiency. More specifically, in this contribution we highlighted our work on TiO <span class="hlt">2</span> anatase mesocrystals of colloidal size. These sophisticate materials; showing excellent textural properties, have remarkable electrochemical performance as anodes for Li/<span class="hlt">Na</span>-ion batteries, with conventional alkyl carbonates electrolytes and safe electrolytes based on ionic liquids. © 2018 The Chemical Society of Japan & Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018MMTB..tmp..897L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018MMTB..tmp..897L"><span>Mechanism of <span class="hlt">Na</span><span class="hlt">2</span>SO4 Promoting Nickel Extraction from Sulfide Concentrates by Sulfation Roasting-Water Leaching</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Li, Guangshi; Cheng, Hongwei; Chen, Sha; Lu, Xionggang; Xu, Qian; Lu, Changyuan</p> <p>2018-04-01</p> <p>As a more environmentally friendly and energy-efficient route, the sulfation roasting-water leaching technique has been developed for highly effective extraction of non-ferrous metals from nickel sulfide concentrate in the presence of a <span class="hlt">Na</span><span class="hlt">2</span>SO4 additive. The effects of several important roasting parameters—the roasting temperature, the addition of <span class="hlt">Na</span><span class="hlt">2</span>SO4, the holding time, and the heating rate in particular—have been investigated. The results suggest that about 90 pct Ni, 92 pct Co, 95 pct Cu, and < 1 pct Fe can be leached from the calcine roasted under the optimum conditions. Furthermore, the behavior and mechanism of the <span class="hlt">Na</span><span class="hlt">2</span>SO4 additive in the roasting process have been well addressed by detailed characterization of the roasted product and leaching residue using quantitative phase analysis (QPA) and energy dispersive spectroscopy (EDS) mapping. The <span class="hlt">Na</span><span class="hlt">2</span>SO4 additive was observed to play a noticeable role in promoting the sulfation degree of valuable metals by forming liquid phases [<span class="hlt">Na</span><span class="hlt">2</span>Me(SO4)<span class="hlt">2</span>] at the outermost layer, which can create a suitable dynamic environment for sulfation. Thus, addition of <span class="hlt">Na</span><span class="hlt">2</span>SO4 might be conducive to an alternative metallurgical process involving complex sulfide ores.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018MMTB...49.1136L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018MMTB...49.1136L"><span>Mechanism of <span class="hlt">Na</span><span class="hlt">2</span>SO4 Promoting Nickel Extraction from Sulfide Concentrates by Sulfation Roasting-Water Leaching</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Li, Guangshi; Cheng, Hongwei; Chen, Sha; Lu, Xionggang; Xu, Qian; Lu, Changyuan</p> <p>2018-06-01</p> <p>As a more environmentally friendly and energy-efficient route, the sulfation roasting-water leaching technique has been developed for highly effective extraction of non-ferrous metals from nickel sulfide concentrate in the presence of a <span class="hlt">Na</span><span class="hlt">2</span>SO4 additive. The effects of several important roasting parameters—the roasting temperature, the addition of <span class="hlt">Na</span><span class="hlt">2</span>SO4, the holding time, and the heating rate in particular—have been investigated. The results suggest that about 90 pct Ni, 92 pct Co, 95 pct Cu, and < 1 pct Fe can be leached from the calcine roasted under the optimum conditions. Furthermore, the behavior and mechanism of the <span class="hlt">Na</span><span class="hlt">2</span>SO4 additive in the roasting process have been well addressed by detailed characterization of the roasted product and leaching residue using quantitative phase analysis (QPA) and energy dispersive spectroscopy (EDS) mapping. The <span class="hlt">Na</span><span class="hlt">2</span>SO4 additive was observed to play a noticeable role in promoting the sulfation degree of valuable metals by forming liquid phases [<span class="hlt">Na</span><span class="hlt">2</span>Me(SO4)<span class="hlt">2</span>] at the outermost layer, which can create a suitable dynamic environment for sulfation. Thus, addition of <span class="hlt">Na</span><span class="hlt">2</span>SO4 might be conducive to an alternative metallurgical process involving complex sulfide ores.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2000SurSc.451..160Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2000SurSc.451..160Y"><span>Desorption induced by electronic transitions of <span class="hlt">Na</span> from SiO<span class="hlt">2</span>: relevance to tenuous planetary atmospheres.</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yakshinskiy, B. V.; Madey, T. E.</p> <p>2000-04-01</p> <p>The authors have studied the desorption induced by electronic transitions (DIET) of <span class="hlt">Na</span> adsorbed on model mineral surfaces, i.e. amorphous, stoichiometric SiO<span class="hlt">2</span> films. They find that electron stimulated desorption (ESD) of atomic <span class="hlt">Na</span> occurs for electron energy thresholds as low as ≡4 eV, that desorption cross-sections are high (≡1×10-19cm<span class="hlt">2</span> at 11 eV), and that desorbing atoms are 'hot', with suprathermal velocities. The estimated <span class="hlt">Na</span> desorption rate from the lunar surface via ESD by solar wind electrons is a small fraction of the rate needed to sustain the <span class="hlt">Na</span> atmosphere. However, the solar photon flux at energies ≥5 eV exceeds the solar wind electron flux by orders of magnitude; there are sufficient ultraviolet photons incident on the lunar surface to contribute substantially to the lunar <span class="hlt">Na</span> atmosphere via PSD of <span class="hlt">Na</span> from the surface.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018SSCom.271...29R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018SSCom.271...29R"><span>Half-metallicity in new Heusler alloys <span class="hlt">Na</span>TO<span class="hlt">2</span> (T=Sc, Ti, V, Cr, and Mn): 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>Rajabi, Kh; Ahmadian, F.</p> <p>2018-03-01</p> <p>On the basis of the full-potential linearized augmented plane wave (FPLAPW) method within density functional theory (DFT), electronic structure and magnetic properties of Heusler alloys <span class="hlt">Na</span>TO<span class="hlt">2</span> (T = Sc, Ti, V, Cr, and Mn) were investigated. The negative values of formation energy showed that these compounds can be experimentally synthesized. Results showed that in all compounds, AlCu<span class="hlt">2</span>Mn-type structure was the most favorable one. The <span class="hlt">Na</span>TO<span class="hlt">2</span> (T = Sc, Ti, V, Cr, and Mn) alloys were HM ferromagnets except <span class="hlt">Na</span>ScO<span class="hlt">2</span> (in both structures which were nonmagnetic semiconductors) and <span class="hlt">Na</span>VO<span class="hlt">2</span> (in AlCu<span class="hlt">2</span>Mn-type structure which was a magnetic semiconductor). The origin of half-metallicity was also verified in HM alloys. <span class="hlt">Na</span>CrO<span class="hlt">2</span> and <span class="hlt">Na</span>VO<span class="hlt">2</span> alloys had higher half-metallic band gaps in comparison with Heusler alloys including and excluding transition metals. The total magnetic moments of HM <span class="hlt">Na</span>TO<span class="hlt">2</span> (T = Ti, V, Cr, and Mn) alloys obeyed Slater-Pauling rule (Mtot = Ztot-12). Among <span class="hlt">Na</span>TO<span class="hlt">2</span> (T = Sc, Ti, V, Cr, and Mn) alloys, <span class="hlt">Na</span>CrO<span class="hlt">2</span> had the highest robustness of half-metallicity with variation of lattice constant in both structures.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/12194529','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/12194529"><span>Microstructural dependence on relevant physical-mechanical properties on SiO<span class="hlt">2</span>-<span class="hlt">Na</span><span class="hlt">2</span>O-CaO-P<span class="hlt">2</span>O5 biological glasses.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Rajendran, V; Begum, A Nishara; Azooz, M A; el Batal, F H</p> <p>2002-11-01</p> <p>Bioactive glasses of the system SiO<span class="hlt">2</span>-<span class="hlt">Na</span><span class="hlt">2</span>O-CaO-P<span class="hlt">2</span>O5 have been prepared by the normal melting and annealing technique. The elastic moduli, attenuation, Vickers hardness, fracture toughness and fracture surface energy have been obtained using the known method at room temperature. The temperature dependence of elastic moduli and attenuation measurements have been extended over a wide range of temperature from 150 to 500 K. The SiO<span class="hlt">2</span> content dependence of velocities, attenuation, elastic moduli, and other parameters show an interesting observation at 45 wt% of SiO<span class="hlt">2</span> by exhibiting an anomalous behaviour. A linear relation is developed for Tg, which explores the influence of <span class="hlt">Na</span><span class="hlt">2</span>O on SiO<span class="hlt">2</span>-<span class="hlt">Na</span><span class="hlt">2</span>O-CaO-P<span class="hlt">2</span>O5 bioactive glasses. The measured hardness, fracture toughness and fracture surface energy show a linear relation with Young's modulus. It is also interesting to note that the observed results are functions of polymerisation and the number of non-bridging oxygens (NBO) prevailing in the network with change in SiO<span class="hlt">2</span> content. The temperature dependence of velocities, attenuation and elastic moduli show the existence of softening in the glass network structure as temperature increases.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3349317','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3349317"><span>Hydrogen Generation from Al-NiCl<span class="hlt">2</span>/<span class="hlt">Na</span>BH4 Mixture Affected by Lanthanum Metal</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Qiang Sun, Wen; Fan, Mei-Qiang; Fei, Yong; Pan, Hua; Wang, Liang Liang; Yao, Jun</p> <p>2012-01-01</p> <p>The effect of La on Al/<span class="hlt">Na</span>BH4 hydrolysis was elaborated in the present paper. Hydrogen generation amount increases but hydrogen generation rate decreases with La content increasing. There is an optimized composition that Al-15 wt% La-5 wt% NiCl<span class="hlt">2</span>/<span class="hlt">Na</span>BH4 mixture (Al-15 wt% La-5 wt% NiCl<span class="hlt">2</span>/<span class="hlt">Na</span>BH4 weight ratio, 1 : 3) has 126 mL g−1 min−1 maximum hydrogen generation rate and 1764 mL g−1 hydrogen generation amount within 60 min. The efficiency is 88%. Combined with NiCl<span class="hlt">2</span>, La has great effect on <span class="hlt">Na</span>BH4 hydrolysis but has little effect on Al hydrolysis. Increasing La content is helpful to decrease the particle size of Al-La-NiCl<span class="hlt">2</span> in the milling process, which induces that the hydrolysis byproduct Ni<span class="hlt">2</span>B is highly distributed into Al(OH)3 and the catalytic reactivity of Ni<span class="hlt">2</span>B/Al(OH)3 is increased therefore. But hydrolysis byproduct La(OH)3 deposits on Al surface and leads to some side effect. The Al-La-NiCl<span class="hlt">2</span>/<span class="hlt">Na</span>BH4 mixture has good stability in low temperature and its hydrolytic performance can be improved with increasing global temperature. Therefore, the mixture has good safety and can be applied as on board hydrogen generation material. PMID:22619596</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22619596','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22619596"><span>Hydrogen generation from Al-NiCl<span class="hlt">2</span>/<span class="hlt">Na</span>BH4 mixture affected by lanthanum metal.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Sun, Wen Qiang; Fan, Mei-Qiang; Fei, Yong; Pan, Hua; Wang, Liang Liang; Yao, Jun</p> <p>2012-01-01</p> <p>The effect of La on Al/<span class="hlt">Na</span>BH(4) hydrolysis was elaborated in the present paper. Hydrogen generation amount increases but hydrogen generation rate decreases with La content increasing. There is an optimized composition that Al-15 wt% La-5 wt% NiCl(<span class="hlt">2</span>)/<span class="hlt">Na</span>BH(4) mixture (Al-15 wt% La-5 wt% NiCl(<span class="hlt">2</span>)/<span class="hlt">Na</span>BH(4) weight ratio, 1 : 3) has 126 mL g(-1 )min(-1) maximum hydrogen generation rate and 1764 mL g(-1) hydrogen generation amount within 60 min. The efficiency is 88%. Combined with NiCl(<span class="hlt">2</span>), La has great effect on <span class="hlt">Na</span>BH(4) hydrolysis but has little effect on Al hydrolysis. Increasing La content is helpful to decrease the particle size of Al-La-NiCl(<span class="hlt">2</span>) in the milling process, which induces that the hydrolysis byproduct Ni(<span class="hlt">2</span>)B is highly distributed into Al(OH)(3) and the catalytic reactivity of Ni(<span class="hlt">2</span>)B/Al(OH)(3) is increased therefore. But hydrolysis byproduct La(OH)(3) deposits on Al surface and leads to some side effect. The Al-La-NiCl(<span class="hlt">2</span>)/<span class="hlt">Na</span>BH(4) mixture has good stability in low temperature and its hydrolytic performance can be improved with increasing global temperature. Therefore, the mixture has good safety and can be applied as on board hydrogen generation material.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017PhRvB..96s5137H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017PhRvB..96s5137H"><span>High-pressure versus isoelectronic doping effect on the honeycomb iridate <span class="hlt">Na</span><span class="hlt">2</span>IrO3</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hermann, V.; Ebad-Allah, J.; Freund, F.; Pietsch, I. M.; Jesche, A.; Tsirlin, A. A.; Deisenhofer, J.; Hanfland, M.; Gegenwart, P.; Kuntscher, C. A.</p> <p>2017-11-01</p> <p>We study the effect of isoelectronic doping and external pressure in tuning the ground state of the honeycomb iridate <span class="hlt">Na</span><span class="hlt">2</span>IrO3 by combining optical spectroscopy with synchrotron x-ray diffraction measurements on single crystals. The obtained optical conductivity of <span class="hlt">Na</span><span class="hlt">2</span>IrO3 is discussed in terms of a Mott-insulating picture versus the formation of quasimolecular orbitals and in terms of Kitaev interactions. With increasing Li content x , (<span class="hlt">Na</span>1 -xLix )<span class="hlt">2</span>IrO3 moves deeper into the Mott-insulating regime, and there are indications that up to a doping level of 24% the compound comes closer to the Kitaev limit. The optical conductivity spectrum of single-crystalline α -Li<span class="hlt">2</span>IrO3 does not follow the trends observed for the series up to x =0.24 . There are strong indications that α -Li<span class="hlt">2</span>IrO3 is not as close to the Kitaev limit as <span class="hlt">Na</span><span class="hlt">2</span>IrO3 and lies closer to the quasimolecular orbital picture instead. Except for the pressure-induced hardening of the phonon modes, the optical properties of <span class="hlt">Na</span><span class="hlt">2</span>IrO3 seem to be robust against external pressure. Possible explanations of the unexpected evolution of the optical conductivity with isolectronic doping and the drastic change between x =0.24 and x =1 are given by comparing the pressure-induced changes of lattice parameters and the optical conductivity with the corresponding changes induced by doping.</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..MARK22014D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016APS..MARK22014D"><span>Doping Li and K into <span class="hlt">Na</span><span class="hlt">2</span>ZrO3 Sorbent to Improve Its CO<span class="hlt">2</span> Capture Capability</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Duan, Yuhua</p> <p></p> <p>Carbon dioxide is one of the major combustion products which once released into the air can contribute to global climate change. Solid sorbents have been reported in several previous studies to be promising candidates for CO<span class="hlt">2</span> sorbent applications due to their high CO<span class="hlt">2</span> absorption capacities at moderate working temperatures. However, at a given CO<span class="hlt">2</span> pressure, the turnover temperature (Tt) of an individual solid capture CO<span class="hlt">2</span> reaction is fixed and may be outside the operating temperature range (ΔTo) for a particularly capture technology. In order to shift such Tt for a solid into the range of ΔTo, its corresponding thermodynamic property must be changed by changing its structure by reacting (mixing) with other materials or doping with other elements. As an example, by combining thermodynamic database searching with ab initio thermodynamics calculations, in this work, we explored the Li- and K-doping effects on the Tt shifts of <span class="hlt">Na</span><span class="hlt">2</span>ZrO3 at different doping levels. The obtained results showed that compared to pure <span class="hlt">Na</span><span class="hlt">2</span>ZrO3, the Li- and K-doped mixtures <span class="hlt">Na</span><span class="hlt">2</span>-αMαZrO3 (M =Li, K) have lower Tt and higher CO<span class="hlt">2</span> capture capacities.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JPS...365..339M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JPS...365..339M"><span>Elucidation of reaction mechanisms of Ni<span class="hlt">2</span>SnP in Li-ion and <span class="hlt">Na</span>-ion systems</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Marino, C.; Dupré, N.; Villevieille, C.</p> <p>2017-10-01</p> <p>Electrochemical performance of Ni<span class="hlt">2</span>SnP was assessed in Li-ion and <span class="hlt">Na</span>-ion battery systems. When cycled versus Li, Ni<span class="hlt">2</span>SnP exhibited a reversible specific charge of 700 mAh.g-1 (theoretical specific charge: 742 mAh.g-1). In the <span class="hlt">Na</span> system, the specific observed charge was ca. 200 mAh.g-1 (theoretical specific charge: 676 mAh.g-1). X-ray diffraction, Ni K-edge X-ray absorption spectroscopy, and 31P and 7Li/23<span class="hlt">Na</span> nuclear magnetic resonance spectroscopy were used to elucidate the electrochemical mechanisms in both systems. Versus Li, Ni<span class="hlt">2</span>SnP undergoes a conversion reaction resulting in the extrusion of Ni and the alloying of Li-Sn and Li-P. On delithiation, the material partially recombines into a Sn- and Ni-deficient form. In the <span class="hlt">Na</span> system, Ni<span class="hlt">2</span>SnP reacts through the conversion of P into <span class="hlt">Na</span>3P. These results indicate that the recombination of the pristine material (even partially) increases cycling stability.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2007M%26PS...42.1221F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2007M%26PS...42.1221F"><span>Al-Mg isotopic evidence for episodic alteration of Ca-Al-rich inclusions from Allende</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Fagan, T. J.; Guan, Y.; MacPherson, G. J.</p> <p>2007-08-01</p> <p>Textures, mineral assemblages, and Al-Mg isotope systematics indicate a protracted, episodic secondary mineralization history for Allende Ca-Al-rich inclusions (<span class="hlt">CAIs</span>). Detailed observations from one type B1 <span class="hlt">CAI</span>, one B<span class="hlt">2</span>, one compact type A (CTA), and one fluffy type A (FTA) indicate that these diverse types of <span class="hlt">CAIs</span> are characterized by two distinct textural and mineralogic types of secondary mineralization: (1) grossular-rich domains, concentrated along melilite grain boundaries in <span class="hlt">CAI</span> interiors, and (<span class="hlt">2</span>) feldspathoid-bearing domains, confined mostly to <span class="hlt">CAI</span> margins just interior to the Wark-Lovering rim sequence. The Al-Mg isotopic compositions of most secondary minerals in the type B1 <span class="hlt">CAI</span>, and some secondary minerals in the other <span class="hlt">CAIs</span>, show no resolvable excesses of 26Mg, whereas the primary <span class="hlt">CAI</span> phases mostly yield correlated excesses of 26Mg with increasing Al/Mg corresponding to "canonical" initial 26Al/27Al ˜ 4.5-5 × 10-5. These secondary minerals formed at least 3 Ma after the primary <span class="hlt">CAI</span> minerals. All but two analyses of secondary minerals from the fluffy type-A <span class="hlt">CAI</span> define a correlated increase in 26Mg/24Mg with increasing Al/Mg, yielding (26Al/27Al)0 = (4.9 ± <span class="hlt">2</span>.8) × 10-6. The secondary minerals in this <span class="hlt">CAI</span> formed 1.8-3.<span class="hlt">2</span> Ma after the primary <span class="hlt">CAI</span> minerals. In both cases, the timing of secondary alteration is consistent with, but does not necessarily require, alteration in an asteroidal setting. One grossular from the type B<span class="hlt">2</span> <span class="hlt">CAI</span>, and several grossular and secondary feldspar analyses from the compact type A <span class="hlt">CAI</span>, have excesses of 26Mg consistent with initial 26Al/27Al ˜ 4.5 × 10-5. Especially in the compact type A <span class="hlt">CAI</span>, where 26Mg/24Mg in grossular correlates with increasing Al/Mg, these 26Mg excesses are almost certainly due to in situ decay of 26Al. They indicate a nebular setting for formation of the grossular. The preservation of these diverse isotopic patterns indicates that heating on the Allende parent body was not pervasive enough to reset isotopic</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018MinPe.112..219K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018MinPe.112..219K"><span>First investigations on the quaternary system <span class="hlt">Na</span><span class="hlt">2</span>O-K<span class="hlt">2</span>O-CaO-SiO<span class="hlt">2</span>: synthesis and crystal structure of the mixed alkali calcium silicate K1.08<span class="hlt">Na</span>0.92Ca6Si4O15</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kahlenberg, Volker; Mayerl, Michael Jean-Philippe; Schmidmair, Daniela; Krüger, Hannes; Tribus, Martina</p> <p>2018-04-01</p> <p>In the course of an exploratory study on the quaternary system <span class="hlt">Na</span><span class="hlt">2</span>O-K<span class="hlt">2</span>O-CaO-SiO<span class="hlt">2</span> single crystals of the first anhydrous sodium potassium calcium silicate have been obtained from slow cooling of a melt in the range between 1250 and 1050 °C. Electron probe micro analysis suggested the following idealized molar ratios of the oxides for the novel compound: K<span class="hlt">2</span>O:<span class="hlt">Na</span><span class="hlt">2</span>O:CaO:SiO<span class="hlt">2</span> = 1:1:12:8 (or KNaCa6Si4O15). Single-crystal diffraction measurements on a crystal with chemical composition K1.08<span class="hlt">Na</span>0.92Ca6Si4O15 resulted in the following basic crystallographic data: monoclinic symmetry, space group P 21/ c, a = 8.9618(9) Å, b = 7.3594(6) Å, c = 11.2453(11) Å, β= 107.54(1)°, V = 707.<span class="hlt">2</span>(1) Å3, Z = <span class="hlt">2</span>. Structure solution was performed using direct methods. The final least-squares refinement converged at a residual of R(|F|) = 0.0346 for 1288 independent reflections and 125 parameters. From a structural point of view, K1.08<span class="hlt">Na</span>0.92Ca6Si4O15 belongs to the group of mixed-anion silicates containing [Si<span class="hlt">2</span>O7]- and [SiO4]-units in the ratio 1:<span class="hlt">2</span>. The mono- and divalent cations occupy a total of four crystallographically independent positions located in voids between the tetrahedra. Three of these sites are exclusively occupied by calcium. The fourth site is occupied by 54(1)% K and 46%(1) <span class="hlt">Na</span>, respectively. Alternatively, the structure can be described as a heteropolyhedral framework based on corner-sharing silicate tetrahedra and [CaO6]-octahedra. The network can build up from kröhnkite-like [Ca(SiO4)<span class="hlt">2</span>O<span class="hlt">2</span>]-chains running along [001]. A detailed comparison with other A<span class="hlt">2</span>B6Si4O15-compounds including topological and group-theoretical aspects is presented.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4866826','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4866826"><span>Cooperative regulation by G proteins and <span class="hlt">Na</span>+ of neuronal GIRK<span class="hlt">2</span> K+ channels</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Wang, Weiwei; Touhara, Kouki K; Weir, Keiko; Bean, Bruce P; MacKinnon, Roderick</p> <p>2016-01-01</p> <p>G protein gated inward rectifier K+ (GIRK) channels open and thereby silence cellular electrical activity when inhibitory G protein coupled receptors (GPCRs) are stimulated. Here we describe an assay to measure neuronal GIRK<span class="hlt">2</span> activity as a function of membrane-anchored G protein concentration. Using this assay we show that four Gβγ subunits bind cooperatively to open GIRK<span class="hlt">2</span>, and that intracellular <span class="hlt">Na</span>+ – which enters neurons during action potentials – further amplifies opening mostly by increasing Gβγ affinity. A <span class="hlt">Na</span>+ amplification function is characterized and used to estimate the concentration of Gβγ subunits that appear in the membrane of mouse dopamine neurons when GABAB receptors are stimulated. We conclude that GIRK<span class="hlt">2</span>, through its dual responsiveness to Gβγ and <span class="hlt">Na</span>+, mediates a form of neuronal inhibition that is amplifiable in the setting of excess electrical activity. DOI: http://dx.doi.org/10.7554/eLife.15751.001 PMID:27074662</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018PCM...tmp...25C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PCM...tmp...25C"><span>Siudaite, <span class="hlt">Na</span>8(Mn<span class="hlt">2</span>+ <span class="hlt">2</span><span class="hlt">Na</span>)Ca6Fe3+ 3Zr3NbSi25O74(OH)<span class="hlt">2</span>Cl·5H<span class="hlt">2</span>O: a new eudialyte-group mineral from the Khibiny alkaline massif, Kola Peninsula</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Chukanov, Nikita V.; Rastsvetaeva, Ramiza K.; Kruszewski, Łukasz; Aksenov, Sergey M.; Rusakov, Vyacheslav S.; Britvin, Sergey N.; Vozchikova, Svetlana A.</p> <p>2018-03-01</p> <p>The new eudialyte-group mineral siudaite, ideally <span class="hlt">Na</span>8(Mn<span class="hlt">2</span>+ <span class="hlt">2</span><span class="hlt">Na</span>)Ca6Fe3+ 3Zr3NbSi25O74(OH)<span class="hlt">2</span>Cl·5H<span class="hlt">2</span>O, was discovered in a peralkaline pegmatite situated at the Eveslogchorr Mt., Khibiny alkaline massif, Kola Peninsula, Russia. The associated minerals are aegirine, albite, microcline, nepheline, astrophyllite, and loparite-(Ce). Siudaite forms yellow to brownish-yellow equant anhedral grains up to 1.5 cm across. Its lustre is vitreous, and the streak is white. Cleavage is none observed. The Mohs' hardness is 4½. Density measured by hydrostatic weighing is <span class="hlt">2</span>.96(1) g/cm3. Density calculated using the empirical formula is equal to <span class="hlt">2</span>.973 g/cm3. Siudaite is nonpleochroic, optically uniaxial, negative, with ω = 1.635(1) and ɛ = 1.626(1) (λ = 589 nm). The IR spectrum is given. The chemical composition of siudaite is (wt%; electron microprobe, H<span class="hlt">2</span>O determined by HCN analysis): <span class="hlt">Na</span><span class="hlt">2</span>O 8.40, K<span class="hlt">2</span>O 0.62, CaO 9.81, La<span class="hlt">2</span>O3 1.03, Ce<span class="hlt">2</span>O3 1.62, Pr<span class="hlt">2</span>O3 0.21, Nd<span class="hlt">2</span>O3 0.29, MnO 6.45, Fe<span class="hlt">2</span>O3 4.51. TiO<span class="hlt">2</span> 0.54, ZrO<span class="hlt">2</span> 11.67, HfO<span class="hlt">2</span> 0.29, Nb<span class="hlt">2</span>O5 <span class="hlt">2</span>.76, SiO<span class="hlt">2</span> 47.20, Cl 0.54, H<span class="hlt">2</span>O 3.5, -O = Cl - 0.12, total 99.32. According to Mössbauer spectroscopy data, all iron is trivalent. The empirical formula (based on 24.5 Si atoms pfu, in accordance with structural data) is [<span class="hlt">Na</span>7.57(H<span class="hlt">2</span>O)1.43]Σ9(Mn1.11<span class="hlt">Na</span>0.88Ce0.31La0.20Nd0.05Pr0.04K0.41)Σ3(H<span class="hlt">2</span>O)1.8(Ca5.46Mn0.54)Σ6(Fe3+ 1.76Mn<span class="hlt">2</span>+ 1.19)Σ<span class="hlt">2</span>.95Nb0.65(Ti0.20Si0.50)Σ0.71(Zr<span class="hlt">2</span>.95Hf0.04Ti0.01)Σ3Si24.00Cl0.47O70(OH)<span class="hlt">2</span>Cl0.47·1.82H<span class="hlt">2</span>O. The crystal structure was determined using single-crystal X-ray diffraction data. The new mineral is trigonal, space group R3m, with a = 14.1885(26) Å, c = 29.831(7) Å, V = 5200.8(23) Å3 and Z = 3. Siudaite is chemically related to georgbarsanovite and is its analogue with Fe3+-dominant M<span class="hlt">2</span> site. The strongest lines of the powder X-ray diffraction pattern [d, Å (I, %) (hkl)] are: 6.38 (60) (-114), 4.29 (55) (-225), 3.389 (47) (131), 3.191 (63) (-228). <span class="hlt">2</span>.963 (100) (4-15), <span class="hlt">2</span>.843 (99) (-444), <span class="hlt">2</span>.577 (49) (3-39). Siudaite is named after the Polish</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22617066-nmr-study-paramagnetic-state-low-dimensional-magnets-licu-sub-sub-nacu-sub-sub','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22617066-nmr-study-paramagnetic-state-low-dimensional-magnets-licu-sub-sub-nacu-sub-sub"><span>NMR study of the paramagnetic state of low-dimensional magnets LiCu{sub <span class="hlt">2</span>}O{sub <span class="hlt">2</span>} and <span class="hlt">Na</span>Cu{sub <span class="hlt">2</span>}O{sub <span class="hlt">2</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>Sadykov, A. F., E-mail: sadykov@imp.uran.ru; Piskunov, Yu. V.; Gerashchenko, A. P.</p> <p></p> <p>A comprehensive NMR study of the magnetic properties of single crystal LiCu{sub <span class="hlt">2</span>}O{sub <span class="hlt">2</span>} (LCO) and <span class="hlt">Na</span>Cu{sub <span class="hlt">2</span>}O{sub <span class="hlt">2</span>} (NCO) is carried out in the paramagnetic region of the compounds for various orientations of single crystals in an external magnetic field. The values of the electric-field gradient (EFG) tensor, as well as the dipole and transferred hyperfine magnetic fields for {sup 63,65}Cu, {sup 7}Li, and {sup 23}<span class="hlt">Na</span> nuclei are determined. The results are compared with the data obtained in previous NMR studies of the magnetically ordered state of LCO/NCO cuprates.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JETPL.tmp...73G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JETPL.tmp...73G"><span>Dimerization in honeycomb <span class="hlt">Na</span><span class="hlt">2</span>RuO3 under pressure: a DFT study</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gazizova, D. D.; Ushakov, A. V.; Streltsov, S. V.</p> <p>2018-04-01</p> <p>The structural properties of <span class="hlt">Na</span><span class="hlt">2</span>RuO3 under pressure are studied using density functional theory within the nonmagnetic generalized gradient approximation (GGA). We found that one may expect a structural transition at ˜3 GPa. This structure at the high-pressure phase is exactly the same as the low-temperature structure of Li<span class="hlt">2</span>RuO3 (at ambient pressure) and is characterized by the P21/m space group. Ru ions form dimers in this phase and one may expect strong modification of the electronic and magnetic properties in <span class="hlt">Na</span><span class="hlt">2</span>RuO3 at pressure higher than 3 GPa.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19970018490','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19970018490"><span>Long Life <span class="hlt">Na</span>/NiCl<span class="hlt">2</span> Cells</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Bugga, Ratnakumar V. (Inventor); Surampudi, Subbarao (Inventor); Halpert, Gerald (Inventor)</p> <p>1996-01-01</p> <p>The premature capacity failure of <span class="hlt">Na</span>/NiCl<span class="hlt">2</span> secondary cells due to agglomeration of nickel particles on the surface of the NiCl<span class="hlt">2</span> cathode is prevented by addition of a minor amount such as 10 percent by weight of a transition metal such as Co, Fe or Mn to the cathode. The chlorides of the transition metals have lower potentials than nickel chloride and chlorinate during charge. A uniform dispersion of the transition metals in the cathodes prevents agglomeration of nickel, maintains morphology of the electrode, maintains the electrochemical area of the electrode and thus maintains capacity of the electrode. The additives do not effect sintering. The addition of sulfur to the liquid catholyte is expected to further reduce agglomeration of nickel in the cathode.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017CryRp..62.1055D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017CryRp..62.1055D"><span>A Novel Coordination Polymer Constructed by Hetero-Metal Ions and <span class="hlt">2</span>,3-Pyridine Dicarboxylic Acid: Synthesis and Structure of [Ni<span class="hlt">Na</span><span class="hlt">2</span>(PDC)<span class="hlt">2</span>(μ-H<span class="hlt">2</span>O)(H<span class="hlt">2</span>O)<span class="hlt">2</span>] n</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Dou, Ming-Yu; Lu, Jing</p> <p>2017-12-01</p> <p>A novel coordination polymer containing hetero-metal ions, [Ni<span class="hlt">Na</span><span class="hlt">2</span>(PDC)<span class="hlt">2</span>(μ-H<span class="hlt">2</span>O)(H<span class="hlt">2</span>O)<span class="hlt">2</span>] n , where PDC is <span class="hlt">2</span>,3-pyridine dicarboxylate ion, has been synthesized. In the structure, the PDC ligand chelates and bridges two Ni(II) and two <span class="hlt">Na</span>(I) centers. Two kinds of metal centers are connected by μ4-PDC and μ<span class="hlt">2</span>-H<span class="hlt">2</span>O to form <span class="hlt">2</span>D coordination layers. Hydrogen bonds between coordination water molecules and carboxylate oxygen atoms further link these <span class="hlt">2</span>D coordination layers to form 3D supramolecular network.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018AIPC.1929b0004L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AIPC.1929b0004L"><span>Fabrication of NIR-responsive <span class="hlt">Na</span>YF4:Yb,Tm/anatase TiO<span class="hlt">2</span> composite aerogel</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Li, Fu-Chih; Kitamoto, Yoshitaka</p> <p>2018-01-01</p> <p>3-dimensional interconnected network structure of TiO<span class="hlt">2</span> aerogel has attracted considerable attention to solve environmental issues due to an advanced oxidation process which uses abundant sunlight for the complete minimization of toxic pollutants. The TiO<span class="hlt">2</span> aerogel with high specific surface area, large pores, and low density has a potential to be used as photocatalyst for air and water purification. Nonetheless, due to the larger band gap, TiO<span class="hlt">2</span> semiconductor photocatalysts possess high oxidizing properties under UV light only which occupies 5% of solar energy. To expand the absorption spectrum of TiO<span class="hlt">2</span> aerogel under solar irradiation, the <span class="hlt">Na</span>YF4:Yb,Tm nanoparticles (NPs) are introduced into the TiO<span class="hlt">2</span> aerogel matrix structure. The morphology and crystal structure of the composite aerogel are investigated by transmission electron microscopy and X-ray diffraction, respectively. The particle size of <span class="hlt">Na</span>YF4:Yb,Tm NPs is approximately 40 nm and the crystallite size of TiO<span class="hlt">2</span> is around 10 nm. In addition, the <span class="hlt">Na</span>YF4:Yb,Tm NPs are enclosed by anatase phase of TiO<span class="hlt">2</span> aerogel. The <span class="hlt">Na</span>YF4:Yb,Tm NPs which exist in the TiO<span class="hlt">2</span> aerogel has a capability of transferring NIR light to UV region.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23618246','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23618246"><span>The TAED/H<span class="hlt">2</span>O<span class="hlt">2</span>/<span class="hlt">Na</span>HCO3 system as an approach to low-temperature and near-neutral pH bleaching of cotton.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Long, Xiaoxia; Xu, Changhai; Du, Jinmei; Fu, Shaohai</p> <p>2013-06-05</p> <p>A low-temperature and near-neutral pH bleaching system was conceived for cotton by incorporating TAED, H<span class="hlt">2</span>O<span class="hlt">2</span> and <span class="hlt">Na</span>HCO3. The TAED/H<span class="hlt">2</span>O<span class="hlt">2</span>/<span class="hlt">Na</span>HCO3 system was investigated and optimized for bleaching of cotton using a central composite design (CCD) combined with response surface methodology (RSM). CCD experimental data were fitted to create a response surface quadratic model (RSQM) describing the degree of whiteness of bleached cotton fabric. Analysis of variance for the RSQM revealed that temperature was the most significant variable, followed by [TAED] and time, while [<span class="hlt">Na</span>HCO3] was insignificant. An effective system was conducted by adding 5.75 g L(-1) TAED together with H<span class="hlt">2</span>O<span class="hlt">2</span> and <span class="hlt">Na</span>HCO3 at a molar ratio of 1:<span class="hlt">2.4:2</span>.8 and applied to bleaching of cotton at 70 °C for 40 min. Compared to a commercial bleaching method, the TAED/H<span class="hlt">2</span>O<span class="hlt">2</span>/<span class="hlt">Na</span>HCO3 system provided cotton with comparable degree of whiteness, slightly inferior water absorbency and acceptable dyeability, but had competitive advantage in protecting cotton from severe chemical damage in bleaching. Copyright © 2013 Elsevier Ltd. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018PCM...tmp..217M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PCM...tmp..217M"><span>Synthesis, structure and properties of blödite-type solid solutions, <span class="hlt">Na</span><span class="hlt">2</span>Co1-x Cu x (SO4)<span class="hlt">2</span>·4H<span class="hlt">2</span>O (0 < x ≤ 0.18), and crystal structure of synthetic kröhnkite, <span class="hlt">Na</span><span class="hlt">2</span>Cu(SO4)<span class="hlt">2</span>·<span class="hlt">2</span>H<span class="hlt">2</span>O</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Marinova, Delyana; Wildner, Manfred; Bancheva, Tsvetelina; Stoyanova, Radostina; Georgiev, Mitko; Stoilova, Donka G.</p> <p>2018-03-01</p> <p>Based on different experimental methods—crystallization processes in aqueous solutions, infrared spectroscopy, single-crystal X-ray diffraction, electron paramagnetic resonance (EPR) and TG-DTA-DSC measurements—it has been established that copper ions are included in sodium cobalt sulfate up to about 18 mol%, thus forming limited solid solutions <span class="hlt">Na</span><span class="hlt">2</span>Co1-x Cu x (SO4)<span class="hlt">2</span>·4H<span class="hlt">2</span>O (0 < x ≤ 0.18) with a blödite-type structure. In contrast, cobalt ions are not able to accept the coordination environment of the copper ions in the strongly distorted Cu(H<span class="hlt">2</span>O)<span class="hlt">2</span>O4 octahedra, thus resulting in the crystallization of Co-free kröhnkite. The solid solutions were characterized by vibrational and EPR spectroscopy. DSC measurements reveal that the copper concentration increase leads to increasing values of the enthalpy of dehydration (ΔH deh) and decreasing values of the enthalpy of formation (ΔH f). The crystal structures of synthetic kröhnkite, <span class="hlt">Na</span><span class="hlt">2</span>Cu(SO4)<span class="hlt">2</span>·<span class="hlt">2</span>H<span class="hlt">2</span>O, as well as of three Cu<span class="hlt">2</span>+-bearing mixed crystals of Co-blödite, <span class="hlt">Na</span><span class="hlt">2</span>Co1-x Cu x (SO4)<span class="hlt">2</span>·4H<span class="hlt">2</span>O with x (Cu) ranging from 0.03 to 0.15, have been investigated from single-crystal X-ray diffraction data. The new data for the structure of synthetic kröhnkite facilitated to clarify structural discrepancies found in the literature for natural kröhnkite samples, traced back to a mix-up of lattice parameters. The crystal structures of Co-dominant <span class="hlt">Na</span><span class="hlt">2</span>Co1-x Cu x (SO4)<span class="hlt">2</span>·4H<span class="hlt">2</span>O solid solutions reveal a comparatively weak influence of the Jahn-Teller-affected Cu<span class="hlt">2</span>+ guest cations up to the maximum content of x (Cu) = 0.15. The response of the MO<span class="hlt">2</span>(H<span class="hlt">2</span>O)4 octahedral shape by increased bond-length distortion with Cu content is clear cut (but limited), mainly concerning the M-OH<span class="hlt">2</span> bond lengths, whereas other structural units are hardly affected. However, the specific type of imposed distortion seems to play an important role impeding higher Cu/Co replacement ratios.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29397745','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29397745"><span>Highly Durable <span class="hlt">Na</span><span class="hlt">2</span>V6O16·1.63H<span class="hlt">2</span>O Nanowire Cathode for Aqueous Zinc-Ion Battery.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Hu, Ping; Zhu, Ting; Wang, Xuanpeng; Wei, Xiujuan; Yan, Mengyu; Li, Jiantao; Luo, Wen; Yang, Wei; Zhang, Wencui; Zhou, Liang; Zhou, Zhiqiang; Mai, Liqiang</p> <p>2018-03-14</p> <p>Rechargeable aqueous zinc-ion batteries are highly desirable for grid-scale applications due to their low cost and high safety; however, the poor cycling stability hinders their widespread application. Herein, a highly durable zinc-ion battery system with a <span class="hlt">Na</span> <span class="hlt">2</span> V 6 O 16 ·1.63H <span class="hlt">2</span> O nanowire cathode and an aqueous Zn(CF 3 SO 3 ) <span class="hlt">2</span> electrolyte has been developed. The <span class="hlt">Na</span> <span class="hlt">2</span> V 6 O 16 ·1.63H <span class="hlt">2</span> O nanowires deliver a high specific capacity of 352 mAh g -1 at 50 mA g -1 and exhibit a capacity retention of 90% over 6000 cycles at 5000 mA g -1 , which represents the best cycling performance compared with all previous reports. In contrast, the <span class="hlt">Na</span>V 3 O 8 nanowires maintain only 17% of the initial capacity after 4000 cycles at 5000 mA g -1 . A single-nanowire-based zinc-ion battery is assembled, which reveals the intrinsic Zn <span class="hlt">2</span>+ storage mechanism at nanoscale. The remarkable electrochemical performance especially the long-term cycling stability makes <span class="hlt">Na</span> <span class="hlt">2</span> V 6 O 16 ·1.63H <span class="hlt">2</span> O a promising cathode for a low-cost and safe aqueous zinc-ion battery.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018RJPCA..92.1025M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018RJPCA..92.1025M"><span>Calculating Equilibrium Constants in the SnCl<span class="hlt">2</span>-H<span class="hlt">2</span>O-<span class="hlt">Na</span>OH System According to Potentiometric Titration Data</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Maskaeva, L. N.; Fedorova, E. A.; Yusupov, R. A.; Markov, V. F.</p> <p>2018-05-01</p> <p>The potentiometric titration of tin chloride SnCl<span class="hlt">2</span> is performed in the concentration range of 0.00009-1.1 mol/L with a solution of sodium hydroxide <span class="hlt">Na</span>OH. According to potentiometric titration data based on modeling equilibria in the SnCl<span class="hlt">2</span>-H<span class="hlt">2</span>O-<span class="hlt">Na</span>OH system, basic equations are generated for the main processes, and instability constants are calculated for the resulting hydroxo complexes and equilibrium constants of low-soluble tin(II) compounds. The data will be of interest for specialists in the field of theory of solutions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28130126','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28130126"><span>Identification of residues that control Li+ versus <span class="hlt">Na</span>+ dependent Ca<span class="hlt">2</span>+ exchange at the transport site of the mitochondrial NCLX.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Roy, Soumitra; Dey, Kuntal; Hershfinkel, Michal; Ohana, Ehud; Sekler, Israel</p> <p>2017-06-01</p> <p>The <span class="hlt">Na</span> + /Ca <span class="hlt">2</span>+ /Li + exchanger (NCLX) is a member of the <span class="hlt">Na</span> + /Ca <span class="hlt">2</span>+ exchanger family. NCLX is unique in its capacity to transport both <span class="hlt">Na</span> + and Li + , unlike other members, which are <span class="hlt">Na</span> + selective. The major aim of this study was twofold, i.e., to identify NCLX residues that confer Li + or <span class="hlt">Na</span> + selective Ca <span class="hlt">2</span>+ transport and map their putative location on NCLX cation transport site. We combined molecular modeling to map transport site of NCLX with euryarchaeal H + /Ca <span class="hlt">2</span>+ exchanger, CAX_Af, and fluorescence analysis to monitor Li + versus <span class="hlt">Na</span> + dependent mitochondrial Ca <span class="hlt">2</span>+ efflux of transport site mutants of NCLX in permeabilized cells. Mutation of Asn149, Pro152, Asp153, Gly176, Asn467, Ser468, Gly494 and Asn498 partially or strongly abolished mitochondrial Ca <span class="hlt">2</span>+ exchange activity in intact cells. In permeabilized cells, N149A, P152A, D153A, N467Q, S468T and G494S demonstrated normal Li + /Ca <span class="hlt">2</span>+ exchange activity but a reduced <span class="hlt">Na</span> + /Ca <span class="hlt">2</span>+ exchange activity. On the other hand, D471A showed dramatically reduced Li + /Ca <span class="hlt">2</span>+ exchange, but <span class="hlt">Na</span> + /Ca <span class="hlt">2</span>+ exchange activity was unaffected. Finally, simultaneous mutation of four putative Ca <span class="hlt">2</span>+ binding residues was required to completely abolish both <span class="hlt">Na</span> + /Ca <span class="hlt">2</span>+ and Li + /Ca <span class="hlt">2</span>+ exchange activities. We identified distinct <span class="hlt">Na</span> + and Li + selective residues in the NCLX transport site. We propose that functional segregation in Li + and <span class="hlt">Na</span> + sites reflects the functional properties of NCLX required for Ca <span class="hlt">2</span>+ exchange under the unique membrane potential and ion gradient across the inner mitochondrial membrane. The results of this study provide functional insights into the unique Li + and <span class="hlt">Na</span> + selectivity of the mitochondrial exchanger. This article is part of a Special Issue entitled: ECS Meeting edited by Claus Heizmann, Joachim Krebs and Jacques Haiech. Copyright © 2017 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JPS...356...80P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JPS...356...80P"><span>P<span class="hlt">2</span>-type <span class="hlt">Na</span><span class="hlt">2</span>/3Mn1-xAlxO<span class="hlt">2</span> cathode material for sodium-ion batteries: Al-doped enhanced electrochemical properties and studies on the electrode kinetics</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Pang, Wei-Lin; Zhang, Xiao-Hua; Guo, Jin-Zhi; Li, Jin-Yue; Yan, Xin; Hou, Bao-Hua; Guan, Hong-Yu; Wu, Xing-Long</p> <p>2017-07-01</p> <p>Recently, sodium-ion batteries (SIBs) have been considered as the promising alternative for lithium-ion batteries. Although layered P<span class="hlt">2</span>-type transition metal oxides are an important class of cathode materials for SIBs, there are still some hurdles for the practical applications, including low specific capacity as well as poor cycling and rate properties. In this study, the electrochemical properties of layered Mn-based oxides have been effectively improved via Al doping, which cannot only promote the formation of layered P<span class="hlt">2</span>-type structure in the preparation processes but also stabilize the lattice during the successive <span class="hlt">Na</span>-intercalation/deintercalation due to suppression of the Jahn-Teller distortion of Mn3+. Among the as-prepared series of <span class="hlt">Na</span><span class="hlt">2</span>/3Mn1-xAlxO<span class="hlt">2</span> (x = 0, 1/18, 1/9, and <span class="hlt">2</span>/9), <span class="hlt">Na</span><span class="hlt">2</span>/3Mn8/9Al1/9O<span class="hlt">2</span> with x = 1/9 exhibits the optimal doping effect with the best electrochemical properties, in terms of the highest specific capacity of 162.3 mA h g-1 at 0.1 C, the highest rate capability, and the best cycling stability in comparison to the undoped <span class="hlt">Na</span><span class="hlt">2</span>/3MnO<span class="hlt">2</span> and the other two materials with different Al-doped contents. Both cyclic voltammetry at varied scan rates and galvanostatic intermittent titration technique disclose the optimal electrode kinetics (the highest <span class="hlt">Na</span>-diffusion coefficient) of the best <span class="hlt">Na</span><span class="hlt">2</span>/3Mn8/9Al1/9O<span class="hlt">2</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2008GeCoA..72.5128D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2008GeCoA..72.5128D"><span>Coupled phase and aqueous species equilibrium of the H <span class="hlt">2</span>O-CO <span class="hlt">2</span>-<span class="hlt">Na</span>Cl-CaCO 3 system from 0 to 250 °C, 1 to 1000 bar with <span class="hlt">Na</span>Cl concentrations up to saturation of halite</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Duan, Zhenhao; Li, Dedong</p> <p>2008-10-01</p> <p>A model is developed for the calculation of coupled phase and aqueous species equilibrium in the H <span class="hlt">2</span>O-CO <span class="hlt">2</span>-<span class="hlt">Na</span>Cl-CaCO 3 system from 0 to 250 °C, 1 to 1000 bar with <span class="hlt">Na</span>Cl concentrations up to saturation of halite. The vapor-liquid-solid (calcite, halite) equilibrium together with the chemical equilibrium of H +, <span class="hlt">Na</span> +, Ca <span class="hlt">2</span>+, CaHCO3+, Ca(OH) +, OH -, Cl -, HCO3-, CO32-, CO <span class="hlt">2</span>(aq) and CaCO 3(aq) in the aqueous liquid phase as a function of temperature, pressure, <span class="hlt">Na</span>Cl concentrations, CO <span class="hlt">2</span>(aq) concentrations can be calculated, with accuracy close to those of experiments in the stated T- P- m range, hence calcite solubility, CO <span class="hlt">2</span> gas solubility, alkalinity and pH values can be accurately calculated. The merit and advantage of this model is its predictability, the model was generally not constructed by fitting experimental data. One of the focuses of this study is to predict calcite solubility, with accuracy consistent with the works in previous experimental studies. The resulted model reproduces the following: (1) as temperature increases, the calcite solubility decreases. For example, when temperature increases from 273 to 373 K, calcite solubility decreases by about 50%; (<span class="hlt">2</span>) with the increase of pressure, calcite solubility increases. For example, at 373 K changing pressure from 10 to 500 bar may increase calcite solubility by as much as 30%; (3) dissolved CO <span class="hlt">2</span> can increase calcite solubility substantially; (4) increasing concentration of <span class="hlt">Na</span>Cl up to <span class="hlt">2</span> m will increase calcite solubility, but further increasing <span class="hlt">Na</span>Cl solubility beyond <span class="hlt">2</span> m will decrease its solubility. The functionality of pH value, alkalinity, CO <span class="hlt">2</span> gas solubility, and the concentrations of many aqueous species with temperature, pressure and <span class="hlt">Na</span>Cl (aq) concentrations can be found from the application of this model. Online calculation is made available on www.geochem-model.org/models/h<span class="hlt">2</span>o_co<span class="hlt">2</span>_nacl_caco3/calc.php.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29368917','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29368917"><span>Triple Halide Bridges in Chiral MnII<span class="hlt">2</span>MnIII6<span class="hlt">Na</span>I<span class="hlt">2</span> Cages: Structural and Magnetic Characterization.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Mayans, Júlia; Font-Bardia, Mercè; Escuer, Albert</p> <p>2018-02-05</p> <p>A family of decanuclear chiral clusters with a Mn II <span class="hlt">2</span> Mn III 6 <span class="hlt">Na</span> I <span class="hlt">2</span> core have been synthesized from enantiomerically pure Schiff bases. The new systems consist of two Mn II Mn III 3 <span class="hlt">Na</span> I units linked by rare triple chloro or bromo bridges between the divalent Mn cations. Susceptibility measurements point out the weak antiferromagnetic interaction mediated by these kinds of bridges and afford the first magnetic measurements for the (μ-Br) 3 case.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016MMTB...47..804S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016MMTB...47..804S"><span>The Cathodic Behavior of Ti(III) Ion in a <span class="hlt">Na</span>Cl-<span class="hlt">2</span>CsCl Melt</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Song, Yang; Jiao, Shuqiang; Hu, Liwen; Guo, Zhancheng</p> <p>2016-02-01</p> <p>The cathodic behavior of Ti(III) ions in a <span class="hlt">Na</span>Cl-<span class="hlt">2</span>CsCl melt was investigated by cyclic voltammetry, chronopotentiometry, and square wave voltammetry with a tungsten electrode being the working electrode at different temperatures. The results show that the cathodic behavior of Ti(III) ion consists of two irreversible steps: Ti3+ + e = Ti<span class="hlt">2</span>+ and Ti<span class="hlt">2</span>+ + <span class="hlt">2</span> e = Ti. The diffusion coefficient for the Ti(III) ion in the <span class="hlt">Na</span>Cl-<span class="hlt">2</span>CsCl eutectic is 1.26 × 10-5 cm<span class="hlt">2</span> s-1 at 873 K (600 °C), increases to be 5.57 × 10-5 cm<span class="hlt">2</span> s-1 at 948K (675°C), and further rises to 10.8 × 10-5 cm<span class="hlt">2</span> s-1 at 1023 (750 °C). Moreover, galvanostatic electrolysis performed on a titanium electrode further presents the feasibility of electrodepositing metallic titanium in the molten <span class="hlt">Na</span>Cl-<span class="hlt">2</span>CsCl-TiCl3 system.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li class="active"><span>24</span></li> <li><a href="#" onclick='return showDiv("page_25");'>25</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_24 --> <div id="page_25" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li><a href="#" onclick='return showDiv("page_24");'>24</a></li> <li class="active"><span>25</span></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="481"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24825479','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24825479"><span>Investigating the solubility and cytocompatibility of CaO-<span class="hlt">Na</span><span class="hlt">2</span> O-SiO<span class="hlt">2</span> /TiO<span class="hlt">2</span> bioactive glasses.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wren, Anthony W; Coughlan, Aisling; Smith, Courtney M; Hudson, Sarah P; Laffir, Fathima R; Towler, Mark R</p> <p>2015-02-01</p> <p>This study aims to investigate the solubility of a series of titanium (TiO<span class="hlt">2</span> )-containing bioactive glasses and their subsequent effect on cell viability. Five glasses were synthesized in the composition range SiO<span class="hlt">2</span> -<span class="hlt">Na</span><span class="hlt">2</span> O-CaO with 5 mol % of increments TiO<span class="hlt">2</span> substituted for SiO<span class="hlt">2</span> . Glass solubility was investigated with respect to (1) exposed surface area, (<span class="hlt">2</span>) particle size, (3) incubation time, and (4) compositional effects. Ion release profiles showed that sodium (<span class="hlt">Na</span>(+) ) presented high release rates after 1 day and were unchanged between 7 and 14 days. Calcium (Ca(<span class="hlt">2</span>+) ) release presented a significant change at each time period and was also composition dependent, where a reduction in Ca(<span class="hlt">2</span>+) release is observed with an increase in TiO<span class="hlt">2</span> concentration. Silica (Si(4+) ) release did not present any clear trends while no titanium (Ti(4+) ) was released. Cell numbers were found to increase up to 44%, compared to the growing control population, with a reduction in particle size and with the inclusion of TiO<span class="hlt">2</span> in the glass composition. © 2014 Wiley Periodicals, Inc.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3647785','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3647785"><span><span class="hlt">Na</span>7Cr4(P<span class="hlt">2</span>O7)4PO4</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Bourguiba Fakhar, Noura; Zid, Mohamed Faouzi; Driss, Ahmed</p> <p>2013-01-01</p> <p>The title compound, hepta­sodium tetra­chromium(III) tetra­kis­(diphosphate) orthophosphate, was synthesized by solid-state reaction. Its structure is isotypic with that of <span class="hlt">Na</span>7 M 4(P<span class="hlt">2</span>O7)4PO4 (M = In, Al) compounds and is made up from a three-dimensional [(CrP<span class="hlt">2</span>O7)4PO4]7− framework with channels running along [001]. The three <span class="hlt">Na</span>+ cations are located in the voids of the framework. One of the cations is situated on a general position, one is equally disordered around a twofold rotation axis and one is on a fourfold rotoinversion axis. The isolated PO4 tetra­hedron of the anionic framework is also situated on the -4 axis. Structural relationships between the title compound and different diphosphates containing MP<span class="hlt">2</span>O11 units (M = Mo, V) are discussed. PMID:23723751</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22416776-age-dependent-changes-diastolic-ca-sup-na-sup-concentrations-dystrophic-cardiomyopathy-role-ca-sup-entry-ip-sub','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22416776-age-dependent-changes-diastolic-ca-sup-na-sup-concentrations-dystrophic-cardiomyopathy-role-ca-sup-entry-ip-sub"><span>Age-dependent changes in diastolic Ca{sup <span class="hlt">2</span>+} and <span class="hlt">Na</span>{sup +} concentrations in dystrophic cardiomyopathy: Role of Ca{sup <span class="hlt">2</span>+} entry and IP{sub 3}</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>Mijares, Alfredo; Altamirano, Francisco; Kolster, Juan</p> <p>2014-10-03</p> <p>Highlights: • Age-dependent increase in [Ca{sup <span class="hlt">2</span>+}]{sub d} and [<span class="hlt">Na</span>{sup +}]{sub d} in mdx cardiomyocytes. • Gadolinium significantly reduced both [Ca{sup <span class="hlt">2</span>+}]{sub d} and [<span class="hlt">Na</span>{sup +}]{sub d} at all ages. • IP{sub 3}-pathway inhibition reduced cations concentrations in dystrophic cardiomyocytes. - Abstract: Duchenne muscular dystrophy (DMD) is a lethal X-inherited disease caused by dystrophin deficiency. Besides the relatively well characterized skeletal muscle degenerative processes, DMD is also associated with a dilated cardiomyopathy that leads to progressive heart failure at the end of the second decade. The aim of the present study was to characterize the diastolic Ca{sup <span class="hlt">2</span>+} concentration ([Ca{supmore » <span class="hlt">2</span>+}]{sub d}) and diastolic <span class="hlt">Na</span>{sup +} concentration ([<span class="hlt">Na</span>{sup +}]{sub d}) abnormalities in cardiomyocytes isolated from 3-, 6-, 9-, and 12-month old mdx mice using ion-selective microelectrodes. In addition, the contributions of gadolinium (Gd{sup 3+})-sensitive Ca{sup <span class="hlt">2</span>+} entry and inositol triphosphate (IP{sub 3}) signaling pathways in abnormal [Ca{sup <span class="hlt">2</span>+}]{sub d} and [<span class="hlt">Na</span>{sup +}]{sub d} were investigated. Our results showed an age-dependent increase in both [Ca{sup <span class="hlt">2</span>+}]{sub d} and [<span class="hlt">Na</span>{sup +}]{sub d} in dystrophic cardiomyocytes compared to those isolated from age-matched wt mice. Gd{sup 3+} treatment significantly reduced both [Ca{sup <span class="hlt">2</span>+}]{sub d} and [<span class="hlt">Na</span>{sup +}]{sub d} at all ages. In addition, blockade of the IP{sub 3}-pathway with either U-73122 or xestospongin C significantly reduced ion concentrations in dystrophic cardiomyocytes. Co-treatment with U-73122 and Gd{sup 3+} normalized both [Ca{sup <span class="hlt">2</span>+}]{sub d} and [<span class="hlt">Na</span>{sup +}]{sub d} at all ages in dystrophic cardiomyocytes. These data showed that loss of dystrophin in mdx cardiomyocytes produced an age-dependent intracellular Ca{sup <span class="hlt">2</span>+} and <span class="hlt">Na</span>{sup +} overload mediated at least in part by enhanced Ca{sup <span class="hlt">2</span>+} entry through Gd{sup 3+} sensitive transient receptor potential channels (TRPC), and by IP{sub 3</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014APS..MARD24011Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014APS..MARD24011Y"><span>Structural Stability and Electronic Properties of <span class="hlt">Na</span><span class="hlt">2</span>C6O6 for a Rechargeable Sodium-ion Battery</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yamashita, Tomoki; Fujii, Akihiro; Momida, Hiroyoshi; Oguchi, Tamio</p> <p>2014-03-01</p> <p>Sodium-ion batteries have been explored as a promising alternative to lithium-ion batteries owing to a significant advantage of a natural abundance of sodium. Recently, it has been reported that disodium rhodizonate, <span class="hlt">Na</span><span class="hlt">2</span>C6O6, exhibit good electrochemical properties and cycle performance as a minor-metal free organic cathode for sodium-ion batteries. However, its crystal structures during discharge/charge cycle still remain unclear. In this work, we theoretically propose feasible crystal structures of <span class="hlt">Na</span><span class="hlt">2</span>+xC6O6 using first principles calculations. A structural phase transition has been found: <span class="hlt">Na</span>4C6O6 has a different C6O6 packing arrangement from <span class="hlt">Na</span><span class="hlt">2</span>C6O6. Electronic structures of <span class="hlt">Na</span><span class="hlt">2</span>+xC6O6 during discharge/charge cycle are also discussed. Our predictions could be the key to understanding the discharge/charge process of <span class="hlt">Na</span><span class="hlt">2</span>C6O6. Supported by MEXT program ``Elements Strategy Initiative to Form Core Rersearch Center'' (since 2012), MEXT; Ministry of Education Culture, Sports, Science and Technology, Japan.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/6992516','DOE-PATENT-XML'); return false;" href="https://www.osti.gov/biblio/6992516"><span>Thermochemical generation of hydrogen and oxygen from water. [<span class="hlt">Na</span>MnO/sub <span class="hlt">2</span>/ and TiO/sub <span class="hlt">2</span>/</span></a></p> <p><a target="_blank" href="http://www.osti.gov/doepatents">DOEpatents</a></p> <p>Robinson, P.R.; Bamberger, C.E.</p> <p>1980-02-08</p> <p>A thermochemical cyclic process for the production of hydrogen exploits the reaction between sodium manganate (<span class="hlt">Na</span>MnO/sub <span class="hlt">2</span>/) and titanium dioxide (TiO/sub <span class="hlt">2</span>/) to form sodium titanate (<span class="hlt">Na</span>/sub <span class="hlt">2</span>/TiO/sub 3/), manganese (II) titanate (MnTiO/sub 3/) and oxygen. The titanate mixture is treated with sodium hydroxide, in the presence of steam, to form sodium titanate, sodium manganate (III), water and hydrogen. The sodium titanate-manganate (III) mixture is treated with water to form sodium manganate (III), titanium dioxide and sodium hydroxide. Sodium manganate (III) and titanium dioxide are recycled following dissolution of sodium hydroxide in water.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29606799','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29606799"><span>Arrhenius Behavior of the Bulk <span class="hlt">Na</span>-Ion Conductivity in <span class="hlt">Na</span>3Sc<span class="hlt">2</span>(PO4)3 Single Crystals Observed by Microcontact Impedance Spectroscopy.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Rettenwander, Daniel; Redhammer, Günther J; Guin, Marie; Benisek, Artur; Krüger, Hannes; Guillon, Olivier; Wilkening, Martin; Tietz, Frank; Fleig, Jürgen</p> <p>2018-03-13</p> <p>NASICON-based solid electrolytes with exceptionally high <span class="hlt">Na</span>-ion conductivities are considered to enable future all solid-state <span class="hlt">Na</span>-ion battery technologies. Despite 40 years of research the interrelation between crystal structure and <span class="hlt">Na</span>-ion conduction is still controversially discussed and far from being fully understood. In this study, microcontact impedance spectroscopy combined with single crystal X-ray diffraction, and differential scanning calorimetry is applied to tackle the question how bulk <span class="hlt">Na</span>-ion conductivity σ bulk of sub-mm-sized flux grown <span class="hlt">Na</span> 3 Sc <span class="hlt">2</span> (PO 4 ) 3 (NSP) single crystals is influenced by supposed phase changes (α, β, and γ phase) discussed in literature. Although we found a smooth structural change at around 140 °C, which we assign to the β → γ phase transition, our conductivity data follow a single Arrhenius law from room temperature (RT) up to 220 °C. Obviously, the structural change, being mainly related to decreasing <span class="hlt">Na</span>-ion ordering with increasing temperature, does not cause any jumps in <span class="hlt">Na</span>-ion conductivity or any discontinuities in activation energies E a . Bulk ion dynamics in NSP have so far rarely been documented; here, under ambient conditions, σ bulk turned out to be as high as 3 × 10 -4 S cm -1  at RT ( E a, bulk = 0.39 eV) when directly measured with microcontacts for individual small single crystals.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5871336','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5871336"><span>Arrhenius Behavior of the Bulk <span class="hlt">Na</span>-Ion Conductivity in <span class="hlt">Na</span>3Sc<span class="hlt">2</span>(PO4)3 Single Crystals Observed by Microcontact Impedance Spectroscopy</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p></p> <p>2018-01-01</p> <p>NASICON-based solid electrolytes with exceptionally high <span class="hlt">Na</span>-ion conductivities are considered to enable future all solid-state <span class="hlt">Na</span>-ion battery technologies. Despite 40 years of research the interrelation between crystal structure and <span class="hlt">Na</span>-ion conduction is still controversially discussed and far from being fully understood. In this study, microcontact impedance spectroscopy combined with single crystal X-ray diffraction, and differential scanning calorimetry is applied to tackle the question how bulk <span class="hlt">Na</span>-ion conductivity σbulk of sub-mm-sized flux grown <span class="hlt">Na</span>3Sc<span class="hlt">2</span>(PO4)3 (NSP) single crystals is influenced by supposed phase changes (α, β, and γ phase) discussed in literature. Although we found a smooth structural change at around 140 °C, which we assign to the β → γ phase transition, our conductivity data follow a single Arrhenius law from room temperature (RT) up to 220 °C. Obviously, the structural change, being mainly related to decreasing <span class="hlt">Na</span>-ion ordering with increasing temperature, does not cause any jumps in <span class="hlt">Na</span>-ion conductivity or any discontinuities in activation energies Ea. Bulk ion dynamics in NSP have so far rarely been documented; here, under ambient conditions, σbulk turned out to be as high as 3 × 10–4 S cm–1 at RT (Ea, bulk = 0.39 eV) when directly measured with microcontacts for individual small single crystals. PMID:29606799</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/17415556','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/17415556"><span>Modulation of the reaction cycle of the <span class="hlt">Na</span>+:Ca<span class="hlt">2</span>+, K+ exchanger.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Vedovato, Natascia; Rispoli, Giorgio</p> <p>2007-09-01</p> <p>Ca(<span class="hlt">2</span>+) concentration in retinal photoreceptor rod outer segment (OS) strongly affects the generator potential kinetics and the receptor light adaptation. The response to intense light stimuli delivered in the dark produce potential changes exceeding 40 mV: since the Ca(<span class="hlt">2</span>+) extrusion in the OS is entirely controlled by the <span class="hlt">Na</span>(+):Ca(<span class="hlt">2</span>+), K(+) exchanger, it is important to assess how the exchanger ion transport rate is affected by the voltage and, in general, by intracellular factors. It is indeed known that the cardiac <span class="hlt">Na</span>(+):Ca(<span class="hlt">2</span>+) exchanger is regulated by Mg-ATP via a still unknown metabolic pathway. In the present work, the <span class="hlt">Na</span>(+):Ca(<span class="hlt">2</span>+), K(+) exchanger regulation was investigated in isolated OS, recorded in whole-cell configuration, using ionic conditions that activated maximally the exchanger in both forward and reverse mode. In all species examined (amphibia: Rana esculenta and Ambystoma mexicanum; reptilia: Gecko gecko), the forward (reverse) exchange current increased about linearly for negative (positive) voltages and exhibited outward (inward) rectification for positive (negative) voltages. Since hyperpolarisation increases Ca(<span class="hlt">2</span>+) extrusion rate, the recovery of the dark level of Ca(<span class="hlt">2</span>+) (and, in turn, of the generator potential) after intense light stimuli results accelerated. Mg-ATP increased the size of forward and reverse exchange current by a factor of approximately <span class="hlt">2</span>.3 and approximately <span class="hlt">2</span>.6, respectively, without modifying their voltage dependence. This indicates that Mg-ATP regulates the number of active exchanger sites and/or the exchanger turnover number, although via an unknown mechanism.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19960008848','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19960008848"><span>Growth of binary organic NLO crystals: m.<span class="hlt">NA-p.NA</span> and m.<span class="hlt">NA</span>-CNA system</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Singh, N. B.; Henningsen, T.; Hopkins, R. H.; Mazelsky, R.</p> <p>1993-01-01</p> <p>Experiments were carried out to grow 3.Nitroaniline (m.<span class="hlt">NA</span>) crystals doped with 4.Nitroaniline (p.<span class="hlt">NA</span>) and <span class="hlt">2</span>.chloro 4.Nitroaniline (CNA). The measured undercooling for m.<span class="hlt">NA</span>, p.<span class="hlt">NA</span>, and CNA were 0.21 tm K, 0.23 tm K, and 0.35 tm K respectively, where tm represents the melting temperature of the pure component. Because of the crystals' large heat of fusion and large undercooling, it was not possible to grow good quality crystals with low thermal gradients. In the conventional two-zone Bridgman furnace we had to raise the temperature of the hot zone above the decomposition temperature of CNA, p.<span class="hlt">NA</span>, and m.<span class="hlt">NA</span> to achieve the desired thermal gradient. To avoid decomposition, we used an unconventional Bridgman furnace. Two immiscible liquids, silicone oil and ethylene glycol, were used to build a special two-zone Bridgman furnace. A temperature gradient of 18 K/cm was achieved without exceeding the decomposition temperature of the crystal. The binary crystals, m.<span class="hlt">NA-p.NA</span> and m.<span class="hlt">NA</span>-CNA, were grown in centimeter size in this furnace. X-ray and optical characterization showed good optical quality.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018SurSc.675...47F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018SurSc.675...47F"><span>Oxidation of MnO(100) and <span class="hlt">Na</span>MnO<span class="hlt">2</span> formation: Characterization of Mn<span class="hlt">2</span>+ and Mn3+ surfaces via XPS and water TPD</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Feng, Xu; Cox, David F.</p> <p>2018-09-01</p> <p>The oxidation of clean and <span class="hlt">Na</span> precovered MnO(100) has been investigated by X-ray photoelectron spectroscopy (XPS), low energy electron diffraction (LEED) and temperature programmed desorption (TPD) of adsorbed water. XPS results indicate that Mn3O4-like and Mn<span class="hlt">2</span>O3-like surfaces can be formed by various oxidation treatments of clean and nearly-stoichiometric MnO(100), while a <span class="hlt">Na</span>MnO<span class="hlt">2</span>-like surface can be produced by the oxidation of MnO(100) pre-covered with multilayers of metallic <span class="hlt">Na</span>. Water TPD results indicate that water adsorption/desorption is sensitive to the available oxidation states of surface Mn cations, and can be used to distinguish between surfaces exposing Mn<span class="hlt">2</span>+and Mn3+ cations, or a combination of these oxidation states. Carbon dioxide and water TPD results from the <span class="hlt">Na</span>MnO<span class="hlt">2</span>-like surface indicate that pre-adsorbed water blocks the uptake of CO<span class="hlt">2</span>, while water displaces pre-adsorbed CO<span class="hlt">2</span>. No indication of a strong reactive interaction is observed between CO<span class="hlt">2</span>, water and the <span class="hlt">Na</span>MnO<span class="hlt">2</span>-like surface under the conditions of our study.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JEMat..47.2009R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JEMat..47.2009R"><span>Effect of Substituted Ca on the Thermoelectric and Optoelectronic Properties of <span class="hlt">Na</span>Rh<span class="hlt">2</span>O4 Under Pressure</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Rahnamaye Aliabad, H. A.; Hosseini, N.</p> <p>2018-03-01</p> <p>In this paper, we have used the first principle calculations for investigation of the structural, optoelectronic and thermoelectric properties of <span class="hlt">Na</span>Rh<span class="hlt">2</span>O4 compound and substituted with Ca onto the <span class="hlt">Na</span> sites under pressure. The results show that there are two direct band gaps for the <span class="hlt">Na</span>Rh<span class="hlt">2</span>O4 compound and three indirect band gaps for the CaRh<span class="hlt">2</span>O4 compound at the top of the Fermi level. The size of the band gaps increases almost linearly with the increase of the pressure up to 37 GPa. The calculated density of states for the CaRh<span class="hlt">2</span>O4 compound show that the Ca-3 p state plays a key role for enhancement of the thermoelectric figure of merit ( ZT). We found that the static dielectric function value decreases along the x, y and z directions for the CaRh<span class="hlt">2</span>O4 compound with the increase of the pressure while it is constant along the x and y directions for the <span class="hlt">Na</span>Rh<span class="hlt">2</span>O4 compound. The birefringence properties with metallic nature are achieved from the optical spectra. The thermoelectric results show that the maximum peak of the ZT shifts towards the higher value of temperature for the <span class="hlt">Na</span>Rh<span class="hlt">2</span>O4 compound. The Ca substitution onto the <span class="hlt">Na</span> sites in the <span class="hlt">Na</span>Rh<span class="hlt">2</span>O4 compound enhances the ZT value of 0.79 at 250 K.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28639392','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28639392"><span>High-Performance <span class="hlt">2</span>.6 V Aqueous Asymmetric Supercapacitors based on In Situ Formed <span class="hlt">Na</span>0.5 MnO<span class="hlt">2</span> Nanosheet Assembled Nanowall Arrays.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Jabeen, Nawishta; Hussain, Ahmad; Xia, Qiuying; Sun, Shuo; Zhu, Junwu; Xia, Hui</p> <p>2017-08-01</p> <p>The voltage limit for aqueous asymmetric supercapacitors is usually <span class="hlt">2</span> V, which impedes further improvement in energy density. Here, high <span class="hlt">Na</span> content Birnessite <span class="hlt">Na</span> 0.5 MnO <span class="hlt">2</span> nanosheet assembled nanowall arrays are in situ formed on carbon cloth via electrochemical oxidation. It is interesting to find that the electrode potential window for <span class="hlt">Na</span> 0.5 MnO <span class="hlt">2</span> nanowall arrays can be extended to 0-1.3 V (vs Ag/AgCl) with significantly increased specific capacitance up to 366 F g -1 . The extended potential window for the <span class="hlt">Na</span> 0.5 MnO <span class="hlt">2</span> electrode provides the opportunity to further increase the cell voltage of aqueous asymmetric supercapacitors beyond <span class="hlt">2</span> V. To construct the asymmetric supercapacitor, carbon-coated Fe 3 O 4 nanorod arrays are synthesized as the anode and can stably work in a negative potential window of -1.3 to 0 V (vs Ag/AgCl). For the first time, a <span class="hlt">2</span>.6 V aqueous asymmetric supercapacitor is demonstrated by using <span class="hlt">Na</span> 0.5 MnO <span class="hlt">2</span> nanowall arrays as the cathode and carbon-coated Fe 3 O 4 nanorod arrays as the anode. In particular, the <span class="hlt">2</span>.6 V <span class="hlt">Na</span> 0.5 MnO <span class="hlt">2</span> //Fe 3 O 4 @C asymmetric supercapacitor exhibits a large energy density of up to 81 Wh kg -1 as well as excellent rate capability and cycle performance, outperforming previously reported MnO <span class="hlt">2</span> -based supercapacitors. This work provides new opportunities for developing high-voltage aqueous asymmetric supercapacitors with further increased energy density. © 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015PhRvB..92m4412M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015PhRvB..92m4412M"><span>Fragile magnetic order in the honeycomb lattice Iridate <span class="hlt">Na</span><span class="hlt">2</span>IrO3 revealed by magnetic impurity doping</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mehlawat, Kavita; Sharma, G.; Singh, Yogesh</p> <p>2015-10-01</p> <p>We report the structure, magnetic, and thermal property measurements on single-crystalline and polycrystalline samples of the Ru-substituted honeycomb lattice iridate <span class="hlt">Na</span><span class="hlt">2</span>Ir1 -xRuxO3 (x =0 ,0.05 ,0.1 ,0.15 ,0.<span class="hlt">2</span> ,0.3 ,0.5 ) . The evolution of magnetism in <span class="hlt">Na</span><span class="hlt">2</span>Ir1 -xRuxO3 has been studied using dc and ac magnetic susceptibilities and heat-capacity measurements. The parent compound <span class="hlt">Na</span><span class="hlt">2</span>IrO3 is a spin-orbit-driven Mott insulator with magnetic order of reduced moments below TN=15 K . In the Ru-substituted samples the antiferromagnetic long-range state is replaced by a spin-glass-like state even for the smallest substitution suggesting that the magnetic order in <span class="hlt">Na</span><span class="hlt">2</span>IrO3 is extremely fragile. We argue that these behaviors indicate the importance of nearest-neighbor magnetic exchange in the parent <span class="hlt">Na</span><span class="hlt">2</span>IrO3 . Additionally, all samples show insulating electrical transport.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018ApPhL.112r2907L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018ApPhL.112r2907L"><span>The influence of excess K<span class="hlt">2</span>O on the electrical properties of (K,<span class="hlt">Na</span>)1/<span class="hlt">2</span>Bi1/<span class="hlt">2</span>TiO3 ceramics</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Li, Linhao; Li, Ming; Sinclair, Derek C.</p> <p>2018-04-01</p> <p>The solid solution (Kx<span class="hlt">Na</span>0.50-x)Bi0.50TiO3 (KNBT) between <span class="hlt">Na</span>1/<span class="hlt">2</span>Bi1/<span class="hlt">2</span>TiO3 and K1/<span class="hlt">2</span>Bi1/<span class="hlt">2</span>TiO3 (KBT) has been extensively researched as a candidate lead-free piezoelectric material because of its relatively high Curie temperature and good piezoelectric properties, especially near the morphotropic phase boundary (MPB) at x ˜ 0.10 (20 mol. % KBT). Here, we show that low levels of excess K<span class="hlt">2</span>O in the starting compositions, i.e., (Ky+0.03<span class="hlt">Na</span>0.50-y)Bi0.50TiO3.015 (y-series), can significantly change the conduction mechanism and electrical properties compared to a nominally stoichiometric KNBT series (Kx<span class="hlt">Na</span>0.50-x)Bi0.50TiO3 (x-series). Impedance spectroscopy measurements reveal significantly higher bulk conductivity (σb) values for y ≥ 0.10 samples [activation energy (Ea) ≤ 0.95 eV] compared to the corresponding x-series samples which possess bandgap type electronic conduction (Ea ˜ 1.26-1.85 eV). The largest difference in electrical properties occurs close to the MPB composition (20 mol. % KBT) where y = 0.10 ceramics possess σb (at 300 °C) that is 4 orders of magnitude higher than that of x = 0.10 and the oxide-ion transport number in the former is ˜0.70-0.75 compared to <0.05 in the latter (between 600 and 800 °C). The effect of excess K<span class="hlt">2</span>O can be rationalised on the basis of the (K + <span class="hlt">Na</span>):Bi ratio in the starting composition prior to ceramic processing. This demonstrates the electrical properties of KNBT to be sensitive to low levels of A-site nonstoichiometry and indicates that excess K<span class="hlt">2</span>O in KNBT starting compositions to compensate for volatilisation can lead to undesirable high dielectric loss and leakage currents at elevated temperatures.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JCrGr.480...62P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JCrGr.480...62P"><span>Growth and characterization of <span class="hlt">Na</span><span class="hlt">2</span>Mo<span class="hlt">2</span>O7 crystal scintillators for rare event searches</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Pandey, Indra Raj; Kim, H. J.; Kim, Y. D.</p> <p>2017-12-01</p> <p>Disodium dimolybdate (<span class="hlt">Na</span><span class="hlt">2</span>Mo<span class="hlt">2</span>O7) crystals were grown using the Czochralski technique. The thermal characteristics of the compound were analyzed using thermogravimetric analysis (TGA) and differential scanning calorimetry (DSC) measurements. The crystal structure of the grown sample was confirmed using X-ray diffraction (XRD). Luminescence properties were measured at room and low temperatures, using a light emitting diode (LED) source. Very weak luminescence was observed at room temperature; however, the luminescence intensity was enhanced at low temperatures. The crystal's transmittance spectrum was measured for estimating its optical quality and energy band gap. The grown crystal exhibited a luminescence light yield of 55% compared with CaMoO4 crystals at 10 K, when excited by a 280-nm-wavelength LED source, but does not have the drawbacks of radioactive Ca isotopes. These results suggest that at cryogenic temperatures, <span class="hlt">Na</span><span class="hlt">2</span>Mo<span class="hlt">2</span>O7 crystal scintillators are promising for the detection of dark matter and neutrinoless double beta decay of 100Mo.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018SurSc.669..130K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018SurSc.669..130K"><span>Peierls instability as the insulating origin of the <span class="hlt">Na</span>/Si(111)-(3 × 1) surface with a <span class="hlt">Na</span> coverage of <span class="hlt">2</span>/3 monolayers</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kang, Myung Ho; Kwon, Se Gab; Jung, Sung Chul</p> <p>2018-03-01</p> <p>Density functional theory (DFT) calculations are used to investigate the insulating origin of the <span class="hlt">Na</span>/Si(111)-(3 × 1) surface with a <span class="hlt">Na</span> coverage of <span class="hlt">2</span>/3 monolayers. In the coverage definition, one monolayer refers to one <span class="hlt">Na</span> atom per surface Si atom, so this surface contains an odd number of electrons (i.e., three Si dangling-bond electrons plus two <span class="hlt">Na</span> electrons) per 3 × 1 unit cell. Interestingly, this odd-electron surface has been ascribed to a Mott-Hubbard insulator to account for the measured insulating band structure with a gap of about 0.8 eV. Here, we instead propose a Peierls instability as the origin of the experimental band gap. The concept of Peierls instability is fundamental in one-dimensional metal systems but has not been taken into account in previous studies of this surface. Our DFT calculations demonstrate that the linear chain structure of Si dangling bonds in this surface is energetically unstable with respect to a × <span class="hlt">2</span> buckling modulation, and the buckling-induced band gap of 0.79 eV explains well the measured insulating nature.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/1344666-effects-al2o3-b2o3-li2o-na2o-sio2-nepheline-crystallization-hanford-high-level-waste-glasses','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/1344666-effects-al2o3-b2o3-li2o-na2o-sio2-nepheline-crystallization-hanford-high-level-waste-glasses"><span>Effects of Al<span class="hlt">2</span>O3, B<span class="hlt">2</span>O3, Li<span class="hlt">2</span>O, <span class="hlt">Na</span><span class="hlt">2</span>O, and SiO<span class="hlt">2</span> on Nepheline Crystallization in Hanford High Level Waste Glasses</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>Kroll, Jared O.; Vienna, John D.; Schweiger, Michael J.</p> <p>2016-09-15</p> <p>Nepheline (nominally <span class="hlt">Na</span>AlSiO4) formation during slow cooling of high-alumina (25.4 - 34.5 mass% Al<span class="hlt">2</span>O3) Hanford high level waste glasses may significantly reduce product durability. To investigate the effects of composition on nepheline crystallization, 29 compositions were formulated by adjusting Al<span class="hlt">2</span>O3, B<span class="hlt">2</span>O3, Li<span class="hlt">2</span>O, <span class="hlt">Na</span><span class="hlt">2</span>O, and SiO<span class="hlt">2</span> around a baseline glass that precipitated 12 mass% nepheline. Thirteen of these compositions were generated by adjusting one-component-at-a-time, while two or three components were adjusted to produce the other 16 (with all remaining components staying in the same relative proportions). Quantitative X-ray diffraction was used to determine nepheline concentration in each sample. Twenty two glassesmore » precipitated nepheline, two of which also precipitated eucryptite (nominally LiAlSiO4), and one glass formed only eucryptite upon slow cooling. Increasing <span class="hlt">Na</span><span class="hlt">2</span>O and Li<span class="hlt">2</span>O had the strongest effect in promoting nepheline formation. Increasing B<span class="hlt">2</span>O3 inhibited nepheline formation. SiO<span class="hlt">2</span> and Al<span class="hlt">2</span>O3 showed non-linear behavior related to nepheline formation. The composition effects on nepheline formation in these glasses are reported.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JChPh.148q4304W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JChPh.148q4304W"><span>Observation of photoassociation of ultracold sodium and cesium at the asymptote <span class="hlt">Na</span> (3S1/<span class="hlt">2</span>) + Cs (6P1/<span class="hlt">2</span>)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wu, Jizhou; Liu, Wenliang; Wang, Xiaofeng; Ma, Jie; Li, Dan; Sovkov, Vladimir B.; Xiao, Liantuan; Jia, Suotang</p> <p>2018-05-01</p> <p>We report on the production of ultracold heteronuclear <span class="hlt">Na</span>Cs* molecules in a dual-species magneto-optical trap through photoassociation. The electronically excited molecules are formed below the <span class="hlt">Na</span> (3S1/<span class="hlt">2</span>) + Cs (6P1/<span class="hlt">2</span>) dissociation limit. 12 resonance lines are detected using trap-loss spectroscopy based on a highly sensitive modulation technique. The highest observed rovibrational level exhibits clear hyperfine structure, which is detected for the first time. This structure is simulated within a simplified model consisting of 4 coupled levels belonging to the initially unperturbed Hund's case "a" electronic states, which have been explored in our previous work that dealt with the <span class="hlt">Na</span> (3S1/<span class="hlt">2</span>) + Cs (6P3/<span class="hlt">2</span>) asymptote [W. Liu et al., Phys. Rev. A 94, 032518 (2016)].</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27760755','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27760755"><span>Ca<span class="hlt">2</span>+ permeability and <span class="hlt">Na</span>+ conductance in cellular toxicity caused by hyperactive DEG/ENaC channels.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Matthewman, Cristina; Miller-Fleming, Tyne W; Miller, David M; Bianchi, Laura</p> <p>2016-12-01</p> <p>Hyperactivated DEG/ENaC channels cause neuronal death mediated by intracellular Ca <span class="hlt">2</span>+ overload. Mammalian ASIC1a channels and MEC-4(d) neurotoxic channels in Caenorhabditis elegans both conduct <span class="hlt">Na</span> + and Ca <span class="hlt">2</span>+ , raising the possibility that direct Ca <span class="hlt">2</span>+ influx through these channels contributes to intracellular Ca <span class="hlt">2</span>+ overload. However, we showed that the homologous C. elegans DEG/ENaC channel UNC-8(d) is not Ca <span class="hlt">2</span>+ permeable, yet it is neurotoxic, suggesting that <span class="hlt">Na</span> + influx is sufficient to induce cell death. Interestingly, UNC-8(d) shows small currents due to extracellular Ca <span class="hlt">2</span>+ block in the Xenopus oocyte expression system. Thus, MEC-4(d) and UNC-8(d) differ both in current amplitude and Ca <span class="hlt">2</span>+ permeability. Given that these two channels show a striking difference in toxicity, we wondered how <span class="hlt">Na</span> + conductance vs. Ca <span class="hlt">2</span>+ permeability contributes to cell death. To address this question, we built an UNC-8/MEC-4 chimeric channel that retains the calcium permeability of MEC-4 and characterized its properties in Xenopus oocytes. Our data support the hypothesis that for Ca <span class="hlt">2</span>+ -permeable DEG/ENaC channels, both Ca <span class="hlt">2</span>+ permeability and <span class="hlt">Na</span> + conductance contribute to toxicity. However, for Ca <span class="hlt">2</span>+ -impermeable DEG/ENaCs (e.g., UNC-8), our evidence shows that constitutive <span class="hlt">Na</span> + conductance is sufficient to induce toxicity, and that this effect is enhanced as current amplitude increases. Our work further refines the contribution of different channel properties to cellular toxicity induced by hyperactive DEG/ENaC channels. Copyright © 2016 the American Physiological Society.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5206307','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5206307"><span>Ca<span class="hlt">2</span>+ permeability and <span class="hlt">Na</span>+ conductance in cellular toxicity caused by hyperactive DEG/ENaC channels</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Matthewman, Cristina; Miller-Fleming, Tyne W.; Miller, David M.</p> <p>2016-01-01</p> <p>Hyperactivated DEG/ENaC channels cause neuronal death mediated by intracellular Ca<span class="hlt">2</span>+ overload. Mammalian ASIC1a channels and MEC-4(d) neurotoxic channels in Caenorhabditis elegans both conduct <span class="hlt">Na</span>+ and Ca<span class="hlt">2</span>+, raising the possibility that direct Ca<span class="hlt">2</span>+ influx through these channels contributes to intracellular Ca<span class="hlt">2</span>+ overload. However, we showed that the homologous C. elegans DEG/ENaC channel UNC-8(d) is not Ca<span class="hlt">2</span>+ permeable, yet it is neurotoxic, suggesting that <span class="hlt">Na</span>+ influx is sufficient to induce cell death. Interestingly, UNC-8(d) shows small currents due to extracellular Ca<span class="hlt">2</span>+ block in the Xenopus oocyte expression system. Thus, MEC-4(d) and UNC-8(d) differ both in current amplitude and Ca<span class="hlt">2</span>+ permeability. Given that these two channels show a striking difference in toxicity, we wondered how <span class="hlt">Na</span>+ conductance vs. Ca<span class="hlt">2</span>+ permeability contributes to cell death. To address this question, we built an UNC-8/MEC-4 chimeric channel that retains the calcium permeability of MEC-4 and characterized its properties in Xenopus oocytes. Our data support the hypothesis that for Ca<span class="hlt">2</span>+-permeable DEG/ENaC channels, both Ca<span class="hlt">2</span>+ permeability and <span class="hlt">Na</span>+ conductance contribute to toxicity. However, for Ca<span class="hlt">2</span>+-impermeable DEG/ENaCs (e.g., UNC-8), our evidence shows that constitutive <span class="hlt">Na</span>+ conductance is sufficient to induce toxicity, and that this effect is enhanced as current amplitude increases. Our work further refines the contribution of different channel properties to cellular toxicity induced by hyperactive DEG/ENaC channels. PMID:27760755</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>